Computational fluid dynamics is a compelling apparatus for getting stream flow and anticipating how this flow will react to various limiting boundary conditions. With this learning, the focal point of this research is applying computational fluid elements such as CFD to issues dealing with stream flow measurement/estimation in closed conduits such as pipes utilizing differential stream meters like the Venturi meter. After thorough research from the existing literature, it was determined that the convergent angle (CA) of a standard Venturi meter has not been optimized yet. The range given for a standard ASME Venturi CA is 20–22°. More than 50 models were created and run in ANSYS FLUENT, which was used as a CFD tool. Three β-ratios are taken into consideration here, which are 0.4, 0.5 and 0.6. An optimum value of CA, corresponding to each β, is obtained by finding out the best coefficient of discharge (closest to 0.99) for each test value of CA. Another aspect explored in this research is the relationship between Reynolds number and coefficient of discharge. This is done with the integration of ANSYS FLUENT and laboratory results. The results of this study yield a definite value of CA for each β.

  • The focal point of this research is the optimization of the convergent angle of the Venturi meter.

  • More than 50 models were created and run in ANSYS FLUENT, and three β-ratios are taken into consideration here, which are 0.4, 0.5 and 0.6.

  • An optimum value of convergent angle, corresponding to each β, is obtained by finding out the best coefficient of discharge (closest to 0.99) for each test value of convergent angle.

Graphical Abstract

Graphical Abstract
ASME

American Society for Mechanical Engineers

SCDM

Space Claim Design Modeller

β-ratio

Diameter ratio

CA

Convergent Angle of Venturi Meter

DA

Divergent Angle of Venturi Meter

RCA

Recovery Cone Angle/Divergent Angle

Cd

Coefficient of Discharge

CFD

Computational Fluid Dynamics

PDEs

Partial Differential Equations

Re

Reynolds Number

FEA

Finite Element Analysis

Qact

Actual Discharge

Qtheo

Theoretical Discharge

Accurate flow measurement remains as one of the biggest concerns in many industries because variations in flows of the product can cost these organizations considerable profits. With the increase in demand for highly accurate flow measurement meters, the use and application of Venturi meters has increased many-fold. In general, there are different meters used to compute the fluid flow: the turbine-type meter, rotameter, orifice meter, and the Venturi meters are only a few. Mubarok et al. (2020) conducted experimental and CFD study on six pressure differential flow meters, namely concentric orifice, bottom eccentric orifice, segmental orifice, top eccentric orifice, nozzle and Venturi. CFD analysis indicated that nozzle and Venturi flow meters showed the lowest net pressure-drops among all the instruments and are most suitable for use in geothermal industries.

These meters work on their ability to alter/affect a certain physical property of the flowing fluid and then this alteration is accurately measured. The measured change in that physical property is then related to the flow, for example the physical property being pressure-drop in orifice and Venturi meters.

A Venturi meter, which is a differential pressure meter, is a measuring device that is usually used to measure the flow of a fluid in a pipe. We can likewise utilize a Venturi meter to increase the velocity of a liquid in a pipe at a specific point. Bernoulli's Theorem remains as its working principle. The latest Venturi meter, Figure 1, comprises an ISA 1932 nozzle entrance and a cone-like development of convergent angle (also known as entry angle) of no more than 22°. It is expected to work in a tight Reynolds number range of 1.5 × 104 to 2 × 106. The coefficient of flow/discharge is 0.95–0.99 for a Venturi meter. A Venturi meter as shown in Figure 1 has the following parts, arranged in systematic order for proper working. These are:
  • inlet section, called the convergent cone;

  • cylindrical throat; and

  • gradually diverging cone.

Figure 1

Basic model of a Venturi meter.

Figure 1

Basic model of a Venturi meter.

Close modal

Convergent Inlet Cone: This is the region where the cross-section of the inlet pipe reduces to a conical shape for its connectivity with the throat such that its cross-sectional area decreases from start to end. On one side it is attached to the inlet pipe and its other side is attached to the cylindrical throat. As per the manual of ASME (ASME MFC-3M-2004 2004), the angle of convergence is generally 20–22° and its length of flow is 2.7(Dd). Here D is the diameter of the inlet section and d is the diameter of the throat. The converging region is attached to the inlet pipe throat region at the downstream end. Due to the decrease in the cross-sectional area, the fluid accelerates and static pressure decreases. The maximum cone angle of the converging area is limited to avoid the vena contracta so the flow area will be minimum at the throat. The convergent angle is believed to be a function of β-ratio, as well as the Reynolds number. The ratio of throat diameter to the diameter of inlet pipe is often termed the β-ratio. The β-ratio acts as a very important physical parameter in designing a Venturi meter. Any change in Reynolds number or the β-ratio affects the most efficient convergent angle for that specific Venturi. The accuracy of a Venturi for discharge measurement is well established and documented through various research, however, the design of convergent cones and the associated head loss is not. This is the area under study in the current research. From earlier studies on gases, Venturi tubes with a convergent angle of 10.5° are much smoother than Venturi tubes with the standard or higher convergent angle (Reader-Harris et al. 2001; Tukimin et al. 2016; Demir 2021).

Throat Section: This is the central part of the Venturi meter and has least cross-sectional area. The length is generally equal to the diameter of the throat. Normally, the diameter of the throat is 0.25–0.75 times the diameter of the inlet pipe, but mostly it is 0.5 times the diameter of the inlet pipe. Throughout its length, the diameter of the throat remains unchanged. Experiment and CFD analysis can be conducted to determine the large energy losses occurring in the throat region for an inhomogeneous flow condition and these losses can be minimized by taking large diameter-ratios (β) (Kim et al. 2015). For two-phase flow experiments which include an oil–water phase, the Venturi pressure-drop is found to be inversely proportional to the Venturi β (Elobeid et al. 2018).

Divergent/Recovery Cone Section: The diverging section is the last part of this instrument. On one side, it is attached to the throat cylinder and on the other to the outlet pipe. The diameter of this section gradually increases. As per the manual from ASME, the diverging section has an angle of 5–13°. The diverging angle is less than the converging angle and the reason for the small diverging angle is to avoid any flow separation from the walls and prevent the formation of eddies. Earlier research has been done on optimizing the Venturi recovery cone angle to minimize the head loss. Sharp et al. (2018) optimized the ASME Venturi recovery cone angle to minimize head loss. The paper mentioned that in many instances where highly accurate flow measurement and low head loss are required, Venturi flow meters are a viable option. This study used computational fluid dynamics and laboratory data to demonstrate the relationship between recovery cone angle in the classical Venturi meter design and associated head loss. Nithin et al. (2012) used the computational fluid dynamics (CFD) software ANSYS FLUENT as a tool to perform modelling and simulation of a Venturi meter and reported that the results from ANSYS FLUENT are in reasonable agreement with experimental values.

As evident in different studies, a lot of analysis has already optimized the divergent angle of Venturi meters in order to bring the value of Cd closer to 1. However, little information is available on that value of convergent angle (CA) which results in the least drop in pressure and the best Cd. Therefore, this study focuses on optimizing the CA to minimize the losses in pressure and thereby results in greater discharge value. This sequentially will lead to better instrument performance.

ANSYS FLUENT

The software package used in this study is ANSYS FLUENT. Often known as Analysis System, ANSYS is a software package that is widely renowned for its engineering simulations. FLUENT, which is an ANSYS solver, contains wide physical modelling abilities needed to model the flow, turbulence, heat transfer and reactions for different industrial applications. It is based on computational fluid dynamics (CFD). CFD is a way of using applied mathematics, physics and computational software to visualize how a fluid flows – as well as how the gas or liquid affects objects as it flows past. CFD has the Navier–Stokes equations as its main working principle. These equations give a description of how the pressure, velocity, temperature, and density of a moving fluid are related. The very foundation of CFD is the Navier–Stokes equations, which are a set of partial differential equations (PDEs) that describe the flow of fluid. With CFD, the study area/model is subdivided into many control volumes, which can also be called cells. In each of these cells, the Navier–Stokes PDEs can be rewritten in the form of algebraic equations that relate the temperature, velocity, flow rate, pressure, and other variables, such as species concentrations, to the values of these parameters in the neighboring cells. Numerical simulation can then be used to solve these equations, yielding a complete picture of the flow domain, down to the resolution of the grid. By solving the resulting set of equations iteratively, a complete description of the flow throughout the domain is yielded. In Venturi meters, instead of costly experimental methods, a computational model can be prepared in ANSYS FLUENT and the coefficient of discharge can be easily calculated (Tamhakar et al. 2014). Ameresh et al.(2017) analyzed a varying inlet diameter of a Venturi meter of 25, 30 and 35 mm and theoretical calculations were performed to evaluate mass flow rate of air and the results further successfully validated with ANSYS FLUENT. From the previous studies, Sanghani et al. (2016) used CFD to investigate the effect of different geometrical parameters like divergent cone angle, diameter ratio and throat length on pressure drop in a Venturi meter. By varying one parameter and keeping the other three constant, the effect of each parameter was found. The results showed fluctuation in pressure drop with increase in convergent cone angle, divergent cone angle and throat length. Also, a reduction in pressure drop with increase in diameter ratio was observed. Furthermore, the employability of CFD tools can be extended for wet gas flow through a Venturi tube by utilizing the discrete phase mode (DPM) and standard k-ε model (Jing et al. 2019). Further research suggests that cavitating flow in a Venturi can be successfully modelled utilizing the Singhal cavitation model in ANSYS FLUENT (Dastane et al. 2019).

Numerical simulations and experimental measurements have shown good agreement in studying the influence of convergent angles on cavitating Venturi tubes (Shi et al. 2019). Not only have Venturi meter instruments shown reasonable agreement between experimental and numerical simulation results, but Venturi nozzles that are specifically used for gas metering have also revealed good agreement between numerical and experimental results (von Lavante et al. 2000). Several studies have been conducted on Venturi-related systems in recent years such as the analysis performed by Li et al. (2022), which depicts enhanced understanding of the mechanism of bubble break in a Venturi-type bubble generator using a volume-of-fluid method and large eddy simulation. Similar studies such as optimization of the geometry of a vane-type pre-swirl nozzle and wet gas flow measurement have also been successfully performed with CFD (Lee et al. 2021; Weise et al. 2021). CFD also has the potential to analyze non-Newtonian fluid in two-phase flow (flow of water–kerosene) in a Venturi meter and results have suggested that coefficient of discharge increases with increments in velocity of fluid and decreases continual with increments in gas volume fraction, as indicated in Khayat & Afarideh (2021). The pressure recovery phenomenon has also been studied by Caetano & Lima (2021) using CFD tools in a Venturi meter utilizing three different k-ɛ turbulence models. Chen & Chen (2022) combined the CFD method and machine learning technique to forecast the radial maximum wall shear stress of an ultra-high pressure (UHP) water jet nozzle. Das et al. (2022) designed a Venturi-type mixture to modulate the flow of producer and bio-gases in a diesel engine. A 2D CFD-based model to study the influence of different concentrations of solid mass on the characteristics of the flow and cavitation generation has also been studied for the Venturi meter by Shi et al. (2020). Hence a sufficient review of the literature reveals that ANSYS FLUENT can be used as a model for simulating the flows in flow-measuring devices.

Reynolds number

The Reynolds number (Re) is a dimensionless entity that is utilized to identify the kind of stream flow as laminar, transient or turbulent while coursing through a pipe. Re plays an important role when it comes to pipe flow. In this study the effect of Re on the coefficient of discharge (Cd) for the Venturi meter is also studied. The value of Cd is always less than 1, except in extreme cases such as slurry flow when it can be greater than 1 (Shook & Masliyah 1974). It is further said that in turbulent flows, the fluid viscosity is less significant as compared with the laminar case and the velocity profile takes on a much more uniform shape. A range of turbulent Re have been studied here and the results are compared for conclusion. In a Venturi meter, the discharge coefficient decreases rapidly as the Re decreases. Arun et al. (2010) prepared a computational model of a Venturi meter that can be used as an efficient and easy means for predicting discharge coefficients at low Re and reported that the discharge coefficient decreases rapidly as the Re decreases. Figure 2 gives an idealized curve for Re versus coefficient of discharge (Boyes 2010). It is believed that higher Re shows positive results on the coefficient of discharge as depicted by the graph below. The validity of this graph has been further studied using laboratory experiment in this study. However, the existence of only one β-ratio in the laboratory setup is a shortcoming. Yadav et al. (2022) investigated the effect of rate of flow on coefficient of discharge in a Venturi meter using an experimental procedure, whereas the changes in pressure and velocity of flow for the Venturi meter were analyzed using computational techniques in SOLIDWORKS. It was reported that coefficient of discharge increased with increments in rate of flow for the designed Venturi meter, and computational analysis further indicated that there is no backflow and eddy formation for the designed Venturi meter, which validated the accuracy of the design. In CFD analysis, the relation between the Reynolds number and differential producer discharge coefficient has been obtained through solutions to the steady, Reynolds-averaged Navier–Stokes equations (Hollingshead et al. 2011).
Figure 2

Re vs Cd curve (Boyes 2010).

The primary goal of the investigation was to decide the ideal convergent angle required to limit the pressure drop for various Venturi meters with differing β proportions. The goals are elaborated as:

  • to correctly determine the optimum angle of convergence for a standard ASME Venturi meter out of the given range of 19–22°;

  • testing the β-ratios of 0.4, 0.5 and 0.6 for different angles of convergence and finding the ones yielding the best coefficient of discharge and the least pressure drop;

  • finding the variation of coefficient of discharge at different Reynolds numbers for each convergent angle;

  • to use computational fluid dynamics and laboratory data (for β = 0.6) to demonstrate the effect of the angles of the convergent cone on the coefficient of discharge on common Venturi meter designs.

For this study, the numerical model development and laboratory testing were done at the Delhi Technological University Laboratory. The physical models were created in ANSYS SPACECLAIM, then exported to FLUENT 18.1 for meshing and analysis. Each of the models was created in SPACECLAIM with XY as the plane of geometry as shown in Figure 3. The geometries of the laboratory-tested flow meters were constructed to replicate the specified drawings.
Figure 3

Geometry in SCDM.

Figure 3

Geometry in SCDM.

Close modal

The laboratory test was done on the Venturi model available in the laboratory. The Venturi model in the laboratory had a β-ratio of 0.6. The equipment had two different pipe dimensions with one pipe of inlet pipe diameter as 20 mm and the other as 35 mm. The pipe material was galvanized iron. Other dimensions are elaborated in the Procedure section.

Procedure: ANSYS FLUENT

Once the geometry of each model was constructed in SCDM, meshing was applied to determine the points within the model where numerical computations would occur. After testing of multiple meshing schemes, it was decided that the best size function would be curvature so as to capture the conical transitions in the Venturi geometry, while keeping the element size as 0.003 m. The mesh was made even finer by introducing refinement of the mesh for different elements in the geometry. Once the models had been meshed, there were up to 500,000 computational elements in a model.

Defining boundary conditions is one of the most important processes for each of the models created in SCDM. The flow inlet on the 30-mm-diameter upstream pipe was defined as the velocity inlet, while the exit of the flow meter was defined as a pressure outlet. All other faces in the geometry were identified as walls. The velocity inlet boundary condition was applied, so that the measured flow from the laboratory data could be used for comparison purposes. The pressure outlet condition was employed to ensure that with different fixed outlet pressures the differential pressure through the meter would be unaffected.

FLUENT was then used and the case file for the mesh was read. The tet/hybrid cells shown in Figure 4 were converted to polyhedral cells to significantly reduce the computation time normally associated with hybrid cells. Through the completion of many test runs it was found that the best viscous turbulence models for this study were the standard or the realizable k-ɛ models. However, for this study, standard realizable k-ɛ was used as the difference between the two was meagre. This particular model was used for any runs that had a Reynolds number greater than 2,000. The boundary condition for the velocity inlet was kept as 1 m/s, corresponding to a Reynolds number value of around 33,500. The pressure outlet was kept so as to have the gauge pressure at that point equal to zero.
Figure 4

Figure depicting nodes and elements in a mesh.

Figure 4

Figure depicting nodes and elements in a mesh.

Close modal

The pressure velocity coupling used was the simple consistent algorithm. The under-relaxation factors were set to the default values ranging from 1 to 0.7 for all the variables besides the pressure factor, which was set to 0.3. For this study, standard pressure was used, while the second-order upwind method was applied for momentum and the first-order upwind scheme for kinetic energy and the turbulent dissipation rate. Residual monitors were used to determine if a solution converged to a point where the results had very little difference between successive iterations. For the k-ɛ model application, six different residuals were being monitored: continuity, x, y, and z velocities, k, and ɛ. The study aimed to ensure the utmost iterative accuracy by requiring all the residuals to converge to 0.001, before the model runs were complete. The models were initialized with hybrid initialization as the main initialization schemes with the input velocity to ensure that the initial iterations gave plausible results. During the initialization, there were usually at least 100 iterations performed on the computer while the residuals were plotted and monitored for any unusual increases. Most of the solutions converged within these iterations. The values of pressures obtained at the two tap locations were used to compute the values of discharges, which were then compared with the theoretical values of discharge to obtain an estimation of Cd. To notice the effect of Reynolds number on the convergent angle, three Reynolds numbers were selected, and each convergent angle was tested for those three Re values.

Procedure: laboratory

The laboratory arrangement had two sets of pipes with a common collection tank as shown in Figure 5(b). The two types of pipes had control valves for the adjustment of discharge. The procedure included measurement of head difference as observed in the manometer legs and then conversion into discharge with the help of equations, as given in the Introduction section. This is known as the Qtheo. The other value, that is Qact, is found with the help of the volume in the collection tank divided by the time taken to fill that volume. The ratio of these values gives the value of Cd. Specifications of the laboratory setup shown in Figure 5(a) and 5(b) are given below:
  • Diameter of the pipe-line: 3.5 cm

  • Diameter of the Venturi throat: 2.0 cm

  • β-ratio of the Venturi meter: 0.6

  • Dimensions of the collecting tank: 50 cm × 40 cm

  • Dimensions of the sump tank: 96 cm × 40 cm

  • Centrifugal pump: 0.37 kW, 220 V, 4 A

Figure 5

(a) Laboratory arrangement for the two Venturi meters; (b) collecting tank and sump tank setup.

Figure 5

(a) Laboratory arrangement for the two Venturi meters; (b) collecting tank and sump tank setup.

Close modal

As the study included both laboratory and computational experiments, results were found for both after following the methodology followed in section 3. Both the results were in form of pressure drops between the two pressure taps in the Venturi meter. The head difference thus obtained had to be further analyzed in order to obtain theoretical as well as actual discharge values. These discharge values are further used to obtain the coefficient of discharge. This section presents those values and the calculations as well.

ANSYS FLUENT results

The ANSYS computations were done for three β-ratios (0.4, 0.5 and 0.6). For each β-ratio, 21 models, corresponding to different convergent cone angles and variable Reynolds numbers, were made. The Qtheo is computed by simply multiplying the value of the area of the input pipe with the input velocity (Table 1). We have taken a standard 30 mm pipe.

Table 1

Theoretical discharge at different velocities

β-ratioInput ANSYS velocity, V1 (m/s)Inlet Reynolds numberQtheo (m3/s)
0.4, 0.5, 0.6 0.2966 10,000 2.0964 × 10−4 
34,000 7.06858 × 10−4 
2.966 100,000 2.0964 × 10−3 
β-ratioInput ANSYS velocity, V1 (m/s)Inlet Reynolds numberQtheo (m3/s)
0.4, 0.5, 0.6 0.2966 10,000 2.0964 × 10−4 
34,000 7.06858 × 10−4 
2.966 100,000 2.0964 × 10−3 

When β-ratio = 0.4

It is necessary to understand that the value of actual discharge will differ for every β-ratio. Different ratios imply different entry lengths and therefore difference in turbulence and also increase/decrease in frictional losses. These losses are visible in terms of difference in pressure drops for different β values. The values highlighted in Figure 6 show the pressure at two tap locations of the Venturi. These values are then used to compute the Qact, which is in turn used to compute the coefficient of discharge as shown in Table 2. The values in the table below only correspond to the inlet Reynolds number of around 34,000. These values can be shown graphically, as done in Figure 7. The graph can be used to find the angle corresponding to the maximum Cd.
Table 2

Cd for different converging angles of β-ratio = 0.4

Convergent angleΔp (fluent value) (Pa)Corresponding QactCd
19 19,881.9813 0.0723563 0.9769 
19.5 19,186.910 0.071078 0.9974 
20 19,076.4588 0.0708740 0.9943 
20.5 19,231.5325 0.07116153 0.9933 
21 19,213.5295 0.0712089 0.9926 
21.5 19,336.4099 0.0713553 0.9906 
22 19,337.2105 0.07135678 0.9905 
Convergent angleΔp (fluent value) (Pa)Corresponding QactCd
19 19,881.9813 0.0723563 0.9769 
19.5 19,186.910 0.071078 0.9974 
20 19,076.4588 0.0708740 0.9943 
20.5 19,231.5325 0.07116153 0.9933 
21 19,213.5295 0.0712089 0.9926 
21.5 19,336.4099 0.0713553 0.9906 
22 19,337.2105 0.07135678 0.9905 
Figure 6

Pressure at two pressure taps.

Figure 6

Pressure at two pressure taps.

Close modal
Figure 7

Graph showing the optimum convergent angle for β = 0.4.

Figure 7

Graph showing the optimum convergent angle for β = 0.4.

Close modal
The graph corresponding to inlet Reynolds number of 34,000 shows that the value corresponding to the convergent angle of 20° yields the least value of pressure drop or the best value of coefficient of discharge i.e. 0.9973. Table 3 and Figure 8 show all the values of convergent angle at all the Reynolds numbers under consideration. It is necessary to note that on controlling the convergent angle, 4% of the loss in terms of pressure can be avoided, as given by the pressure-drop values in Table 2.
Table 3

Cd for different Reynolds numbers for β = 0.4

Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.9721 
34,000 0.9769 
100,000 0.9787 
19.5 10,000 0.9887 
34,000 0.9944 
100,000 0.9963 
20 10,000 0.9925 
34,000 0.9973 
100,000 0.9992 
20.5 10,000 0.9901 
34,000 0.9933 
100,000 0.9951 
21 10,000 0.9898 
34,000 0.9926 
100,000 0.995 
21.5 10,000 0.9881 
34,000 0.99 
100,000 0.9911 
22 10,000 0.9897 
34,000 0.9905 
100,000 0.9935 
Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.9721 
34,000 0.9769 
100,000 0.9787 
19.5 10,000 0.9887 
34,000 0.9944 
100,000 0.9963 
20 10,000 0.9925 
34,000 0.9973 
100,000 0.9992 
20.5 10,000 0.9901 
34,000 0.9933 
100,000 0.9951 
21 10,000 0.9898 
34,000 0.9926 
100,000 0.995 
21.5 10,000 0.9881 
34,000 0.99 
100,000 0.9911 
22 10,000 0.9897 
34,000 0.9905 
100,000 0.9935 
Figure 8

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.4.

Figure 8

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.4.

Close modal

Table 3 shows the variation of the Cd over a range of Reynolds numbers, and it solidifies the fact that 20° is the optimum angle for β = 0.4 as on increasing the Reynolds number further for that CA, the Cd value further increases.

When β-ratio = 0.5

The above-mentioned procedure for computation of Cd is repeated for this value of β (Table 4). To our surprise, the optimum value for this β occurs at 19.5°, which is ideally out of the range of ASME limits but was tested for this study. This can be made out from the graph in Figure 9.
Table 4

Cd for different converging angles of β ratio = 0.5

Convergent angleΔp (fluent value) (Pa)Corresponding QactCd
19 7,639.2953 0.07142 0.9896 
19.5 7,606.3997 0.07777 0.9917 
20 7,664.8414 0.071546 0.9879 
20.5 7,782.9383 0.072095 0.9804 
21 7,763.7068 0.072006 0.9816 
21.5 7,747.55301 0.0719312 0.9826 
22 7,787.16984 0.072114 0.9801 
Convergent angleΔp (fluent value) (Pa)Corresponding QactCd
19 7,639.2953 0.07142 0.9896 
19.5 7,606.3997 0.07777 0.9917 
20 7,664.8414 0.071546 0.9879 
20.5 7,782.9383 0.072095 0.9804 
21 7,763.7068 0.072006 0.9816 
21.5 7,747.55301 0.0719312 0.9826 
22 7,787.16984 0.072114 0.9801 
Figure 9

Graph showing the optimum convergent angle for β = 0.5.

Figure 9

Graph showing the optimum convergent angle for β = 0.5.

Close modal

The Reynolds number and variation associated with it for β = 0.5 are depicted in Table 5. The trend stays the same here: with increase in the Reynolds number value the corresponding coefficient of discharge also increases.

Table 5

Cd for different Reynolds numbers for β = 0.5

Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.98 
34,000 0.9896 
100,000 0.9899 
19.5 10,000 0.9901 
34,000 0.9917 
100,000 0.9936 
20 10,000 0.9822 
34,000 0.9879 
100,000 0.9896 
20.5 10,000 0.9796 
34,000 0.9804 
100,000 0.9896 
21 10,000 0.9801 
34,000 0.9816 
100,000 0.9888 
21.5 10,000 0.9799 
34,000 0.9826 
100,000 0.9899 
22 10,000 0.9763 
34,000 0.9822 
100,000 0.9888 
Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.98 
34,000 0.9896 
100,000 0.9899 
19.5 10,000 0.9901 
34,000 0.9917 
100,000 0.9936 
20 10,000 0.9822 
34,000 0.9879 
100,000 0.9896 
20.5 10,000 0.9796 
34,000 0.9804 
100,000 0.9896 
21 10,000 0.9801 
34,000 0.9816 
100,000 0.9888 
21.5 10,000 0.9799 
34,000 0.9826 
100,000 0.9899 
22 10,000 0.9763 
34,000 0.9822 
100,000 0.9888 

The graphical representation of the tabular data in Figure 10 shows the trend followed by all the convergent angles. Most of the curves follow the idealized curve given in Figure 2.
Figure 10

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.5.

Figure 10

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.5.

Close modal

When β-ratio = 0.6

As discussed above, the computations for this ratio are also done in the same manner. Cd shows a huge variation for different CAs. Table 6 shows the variation and the pressure drop as well. This is the highest ratio to be tested in this study. The optimum value of CA in this case lies well within the range specified by ASME. The optimum convergent cone angle comes out to be 20.5°, with the Cd value of 0.995 as shown in Figure 11.
Table 6

Cd for different convergent angles of β-ratio = 0.6

Convergent angleΔp (fluent value)Corresponding QactCd
19 3,511.60513 0.0723103 0.97753 
19.5 3,560.33215 0.0728931 0.96967 
20 3,563.11174 0.0729216 0.96930 
20.5 3,381.03648 0.0710340 0.99509 
21 3,397.6590 0.0712084 0.9926 
21.5 3,393.02758 0.071159 0.99337 
22 3,415.7200 0.0713974 0.99030 
Convergent angleΔp (fluent value)Corresponding QactCd
19 3,511.60513 0.0723103 0.97753 
19.5 3,560.33215 0.0728931 0.96967 
20 3,563.11174 0.0729216 0.96930 
20.5 3,381.03648 0.0710340 0.99509 
21 3,397.6590 0.0712084 0.9926 
21.5 3,393.02758 0.071159 0.99337 
22 3,415.7200 0.0713974 0.99030 
Figure 11

Graph showing the optimum convergent angle for β = 0.6.

Figure 11

Graph showing the optimum convergent angle for β = 0.6.

Close modal
The effect of Re on Cd is computed and shown in Table 7 and graphically for different convergent angles in Figure 12.
Table 7

Cd for different Reynolds numbers for β = 0.6

Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.9701 
34,000 0.9775 
100,000 0.981 
19.5 10,000 0.9676 
34,000 0.9697 
100,000 0.9716 
20 10,000 0.9633 
34,000 0.9693 
100,000 0.9723 
20.5 10,000 0.9928 
34,000 0.995 
100,000 0.9972 
21 10,000 0.9901 
34,000 0.9926 
100,000 0.9963 
21.5 10,000 0.9907 
34,000 0.9933 
100,000 0.9954 
22 10,000 0.9889 
34,000 0.99 
100,000 0.991 
Convergent angleReynolds numberCoefficient of discharge
19 10,000 0.9701 
34,000 0.9775 
100,000 0.981 
19.5 10,000 0.9676 
34,000 0.9697 
100,000 0.9716 
20 10,000 0.9633 
34,000 0.9693 
100,000 0.9723 
20.5 10,000 0.9928 
34,000 0.995 
100,000 0.9972 
21 10,000 0.9901 
34,000 0.9926 
100,000 0.9963 
21.5 10,000 0.9907 
34,000 0.9933 
100,000 0.9954 
22 10,000 0.9889 
34,000 0.99 
100,000 0.991 
Figure 12

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.6.

Figure 12

Graphical representation of coefficient of discharge for different Reynolds numbers and convergent angles for β = 0.6.

Close modal

Laboratory results

As mentioned earlier, the meters present in the DTU laboratory have a β-ratio of 0.6. Therefore, this section shows the comparison of laboratory results for the 0.6 β-ratio with the CFD results. As the convergent angle in the laboratory is constant, the comparison shown is for an inlet convergent angle of 21°, with a recovery cone angle of 5°. The parameter varied here is Reynolds number only.

The laboratory tests are limited to a range of Reynolds numbers as discharge can only be increased to a certain extent. Therefore, the range is limited. Table 8 indicates all the values, including values of Cd as well. The values of Cd against Reynolds number given in Table 8 are plotted in Figure 13. It gives a comparison of CFD and laboratory results. It is observed that laboratory results vary a little from the CFD curve.
Table 8

Experimental results for β = 0.6

Inlet pipe size, mmManometer reading
Δh of water (m)Discharge (m3/s) QtheoDischarge (m3/s) QactReynolds number, ReCd
Left limb (cm)Right limb (cm)
35 27.1 25.2 0.239 8.04 × 10−4 9.06 × 10−4 37,032 0.888 
28.3 24 0.541 1.21 × 10−3 1.30 × 10−3 53,406 0.926 
29.6 22.5 0.894 1.53 × 10−3 1.55 × 10−3 62,667 0.985 
30.8 19.6 1.411 1.97 × 10−3 1.98 × 10−3 80,661 0.989 
20 10.6 9.8 0.108 1.76 × 10−4 2.25 × 10−4 16,087 0.782 
11 9.4 0.217 2.50 × 10−4 3.08 × 10−4 22,078 0.811 
11.5 8.8 0.367 3.35 × 10−4 4.01 × 10−3 29,287 0.835 
16.5 4.0 1.706 7.01 × 10−4 7.61 × 10−3 54,386 0.921 
Inlet pipe size, mmManometer reading
Δh of water (m)Discharge (m3/s) QtheoDischarge (m3/s) QactReynolds number, ReCd
Left limb (cm)Right limb (cm)
35 27.1 25.2 0.239 8.04 × 10−4 9.06 × 10−4 37,032 0.888 
28.3 24 0.541 1.21 × 10−3 1.30 × 10−3 53,406 0.926 
29.6 22.5 0.894 1.53 × 10−3 1.55 × 10−3 62,667 0.985 
30.8 19.6 1.411 1.97 × 10−3 1.98 × 10−3 80,661 0.989 
20 10.6 9.8 0.108 1.76 × 10−4 2.25 × 10−4 16,087 0.782 
11 9.4 0.217 2.50 × 10−4 3.08 × 10−4 22,078 0.811 
11.5 8.8 0.367 3.35 × 10−4 4.01 × 10−3 29,287 0.835 
16.5 4.0 1.706 7.01 × 10−4 7.61 × 10−3 54,386 0.921 
Figure 13

Plot comparing laboratory and CFD results.

Figure 13

Plot comparing laboratory and CFD results.

Close modal

This is because the CFD computations took a smooth wall under consideration, whereas the laboratory conditions vary a little. But, in any case, it is safe to say that the change in Cd is comparable in both the cases. It can be observed through the graph that in the range of study, that is, for Reynolds numbers 20,000–100,000, the coefficient of discharge is observed to increase.

This section discusses some key findings of this study.

  • CFD is capable of producing data predicting pressure loss/drop for many different Venturi designs. As indicated in Gowtham Sanjai et al. (2021) the CFD analysis is the best approach based on accuracy and time constraint. We observed that the convergent angle corresponding to the maximum coefficient of discharge also has the least value of pressure drop through the CFD computation in ANSYS. Table 9 below gives the percentage loss in pressure that can be avoided if the optimum CAs, as pointed out in this research, are used. These tables are collectively derived from the pressure-drop values in Tables 2, 4 and 6.

Table 9

Loss in pressure for different β-ratios

β-ratio% Loss of pressure that can be avoided by providing optimum CA
0.4 4.051 
0.5 2.380 
0.6 5.110 
β-ratio% Loss of pressure that can be avoided by providing optimum CA
0.4 4.051 
0.5 2.380 
0.6 5.110 

Therefore, it is inferred that head loss measurements are sensitive to the convergent cone angle. In other words, slight changes in the convergent angle can result in large increase/decrease in pressure loss as depicted in Table 9.

  • The convergent angles are proved to be a function of β-ratio and Reynolds number. The best CA angle corresponding to each β-ratio as found is given in Table 10.

  • To validate the results obtained by CFD, an equation presented by Reader-Harris et al. (2001) is used:
    (1)
Table 10

Optimum CA corresponding to different β-ratios

β-ratioOptimum CA
0.4 19.5° 
0.5 19.5° 
0.6 20.5° 
β-ratioOptimum CA
0.4 19.5° 
0.5 19.5° 
0.6 20.5° 

This equation is used to relate Cwater, that is, the coefficient of discharge of a Venturi tube, having water as the flowing fluid with different β-ratios. This equation has an uncertainty of 0.74% (based on two standard deviations). To check our results, the value of Cd obtained by our optimum CAs were compared with the Cwater (Table 11). The error percentage is calculated and checked to see whether it is within the limit or not. It is seen from Table 11 that the error is even less than 1%, therefore the result is very well validated.

  • The relationship between Reynolds number and Cd is also very well established here with the help of CFD results and validated through the laboratory results. This falls into line with the results of Yadav et al. (2022), where the researchers reported that the coefficient of discharge increased with increments in rate of flow for the designed Venturi meter.

Table 11

Validation of the CFD results

β-ratioComputed CdCwater from Equation (1)Error (%)
0.4 0.9974 0.9927 0.46 
0.5 0.9917 0.9939 0.22 
0.6 0.9950 0.9951 0.018 
β-ratioComputed CdCwater from Equation (1)Error (%)
0.4 0.9974 0.9927 0.46 
0.5 0.9917 0.9939 0.22 
0.6 0.9950 0.9951 0.018 

The current study employs physical and numerical data to determine optimum CAs for different Venturi designs. Three different β-ratios of 0.40, 0.50, and 0.60 were analyzed with constant Venturi wall roughness along with three different Reynolds numbers and multiple CAs. The CFD-based research presents the optimum value of CA for different β-ratios which can be chosen while manufacturing the Venturi instrument. This will prevent pressure loss and help in discharge estimation with an error less than 0.25%. The study also indicates an increasing CA value on increasing the β-ratio. Additionally, a relationship between Cd & Re has been observed through the analysis in CFD and laboratory experiment. The curve obtained through both the data are seen to follow a rising trend. The trend signifies that on increasing the Reynolds number (within the experimental range), the Cd also increases. The validation brings out a good congruence in the obtained CFD results and the research carried out by Reader-Harris et al. (2001).

This study focuses only on certain values of β-ratio, therefore, as a future scope, more rigorous analysis can be done to establish a clearer relationship between the CA and β-ratio. In addition to CFD, machine learning algorithms can be used for shorter computation times and increased prediction accuracy.

The are grateful to the laboratory staff of the Hydraulics Laboratory of Delhi Technological University (DTU) for their cooperation in carrying out the results of the current research.

All the authors confirm contributions to the paper as follows: ZAK initiated the idea of the research. ZAK and NJ made the CFD model of the Venturi meter in ANSYS FLUENT. NJ performed all the experiments on the Venturi meter in the laboratory. ZAK analyzed and interpreted the results regarding the optimization of the convergent angles of the Venturi meter and subsequent losses associated with it and was a major contributor in writing the manuscript. All authors reviewed the results and approved the final version of the manuscript.

No funding was obtained for this study.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Ameresh
H.
,
DepaSandeep, Ravikanth Raju
P.
&
Venkat Reddy
R.
2017
Investigation of mass flow rate in venturimeter using CFD analysis
.
International Journal of Engineering Research and Applications
7
(
12 (Part-7)
),
86
90
.
Arun
N.
,
Sankarasubramanian
M.
&
Kaushik
M.
2010
CFD analysis on discharge coefficient during non-Newtonian flows through orifice meter
.
International Journal of Engineering Science and Technology
2
(
7
),
3151
3164
.
May 2015
.
ASME MFC-3M-2004
2004
Measurement of Fluid Flow in Pipes Using Orifice, Nozzle and Venturi, Part 4
.
American Society of Mechanical Engineers
,
New York, USA
, pp.
68
83
.
Boyes
W.
2010
Instrumentation Reference Book
, 4th edn.
Butterworth-Heinemann, Burlington, MA, USA
.
Caetano
N. M.
&
Lima
L. E. M.
2021
Numerical assessment of the pressure recovery of the turbulent flow in a Venturi-type device
. In:
Vibration Engineering and Technology of Machinery: Proceedings of VETOMAC XV 2019
(J. M. Balthazar, ed.), Springer, Cham, Switzerland, pp. 121–138. https://doi.org/10.1007/978-3-030-60694-7_7
.
Chen
Y.-J.
&
Chen
Z.-S.
2022
A prediction model of wall shear stress for ultra-high-pressure water-jet nozzle based on hybrid BP neural network
.
Engineering Applications of Computational Fluid Mechanics
16
(
1
),
1902
1920
.
https://doi.org/10.1080/19942060.2022.2123404
.
Das
S.
,
Kashyap
D.
,
Kalita
P.
&
Kulkarni
V.
2022
Computational investigation of a venturi-type mixer for gaseous fuel application in a dual-fuel engine
.
International Journal of Ambient Energy
,
doi:10.1080/01430750.2022.2075932
.
Dastane
G. G.
,
Thakkar
H.
,
Shah
R.
,
Perala
S.
,
Raut
J.
&
Pandit
A. B.
2019
Single and multiphase CFD simulations for designing cavitating venturi
.
Chemical Engineering Research and Design
.
149
,
1
12
.
https://doi.org/10.1016/j.cherd.2019.06.036
.
Demir
S.
2021
CFD analysis of mini Venturi-type air flowmeters for lab-scale applications
.
Sigma Journal of Engineering and Natural Sciences
38
(
2
),
603
612
.
2020
.
Elobeid
M. O.
,
Ahmad
A.
,
Al-Sarkhi
A.
,
Alhems
L. M.
,
Shaahid
S. M.
,
Basha
M.
,
Xiao
J. J.
,
Lastra
R.
&
Ejim
C. E.
2018
Pressure drop measurements in venturi meters of different beta ratios for oil–water flow experiments
.
Arabian Journal for Science and Engineering
43
,
6355
6374
.
https://doi.org/10.1007/s13369-017-3019-8
.
Gowtham Sanjai
S.
,
Suresh
V.
,
Bedi
R.
&
Sumanthran
A.
2021
Experimental approach and CFD analysis on flow devices. In: Advances in Fluid Dynamics: Selected Proceedings of ICAFD 2018 (B. Rushi Kumar, R. Sivaraj & J. Prakash, eds), Springer, Singapore, pp. 427–436. https://doi.org/10.1007/978-981-15-4308-1_33
.
Hollingshead
C. L.
,
Johnson
M. C.
,
Barfuss
S. L.
&
Spall
R. E.
2011
Discharge coefficient performance of Venturi, standard concentric orifice plate, V-cone and wedge flow meters at low Reynolds numbers
.
Journal of Petroleum Science and Engineering
78
(
3–4
),
559
566
.
https://doi.org/10.1016/j.petrol.2011.08.008
.
Jing
J.
,
Yuan
Y.
,
Du
S.
,
Yin
X.
&
Yin
R.
2019
A CFD study of wet gas metering over-reading model under high pressure
.
Flow Measurement and Instrumentation
69
,
101608
.
https://doi.org/10.1016/j.flowmeasinst.2019.101608
.
Khayat
O.
&
Afarideh
H.
2021
Numerical investigation of non-Newtonian liquid–gas flow in venturi flow meter using computational fluid dynamics
.
Iranian Journal of Science & Technology, Transactions of Civil Engineering
45
,
393
401
.
https://doi.org/10.1007/s40997-019-00321-z
.
Kim
J. H.
,
Jung
U. H.
,
Kim
S.
,
Yoon
J.-Y.
&
Choi
Y.-S.
2015
Uncertainty analysis of flow rate measurement for multiphase flow using CFD
.
Acta Mechanica Sinica
31
,
698
707
.
https://doi.org/10.1007/s10409-015-0493-7
.
Lee
J.
,
Lee
H.
,
Park
H.
,
Cho
G.
,
Kim
D.
&
Cho
J.
2021
Design optimization of a vane type pre-swirl nozzle
.
Engineering Applications of Computational Fluid Mechanics
15
(
1
),
164
179
.
https://doi.org/10.1080/19942060.2020.1847199
.
Li
Q.
,
Ming
D.
,
Lei
M.
,
Guo
X.
,
Liu
J.
,
Zhu
H.
,
Fang
L.
&
Wang
Z.
2022
Numerical investigation on the coupled mechanisms of bubble breakup in a venturi-type bubble generator
.
Engineering Applications of Computational Fluid Mechanics
16
(
1
),
229
247
.
https://doi.org/10.1080/19942060.2021.2008501
.
Mubarok
M. H.
,
Cater
J. E.
&
Zarrouk
S. J.
2020
Comparative CFD modelling of pressure differential flow meters for measuring two-phase geothermal fluid flow
.
Geothermics
86
,
101801
.
https://doi.org/10.1016/j.geothermics.2020.101801
.
Nithin
T.
,
Jain
N.
&
Hiriyannaiah
A.
2012
Optimization of venturi flow meter model for the angle of divergence with minimal pressure drop by computational fluid dynamics method
. In:
Proceedings of International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM – 2012), Bonfring, Coimbatore, Tamilnadu, India
, pp.
658
670
.
Reader-Harris
M. J.
,
Brunton
W. C.
,
Gibson
J. J.
,
Hodges
D.
&
Nicholson
I. G.
2001
Discharge coefficients of Venturi tubes with standard and non-standard convergent angles
.
Flow Measurement and Instrumentation
12
(
2
),
135
145
.
https://doi.org/10.1016/S0955-5986(01)00007-3
.
Sanghani
C. R.
,
Jayani
D. C.
,
Jadvani
N. R.
,
Dobariya
H. N.
&
Jasoliya
K. R.
2016
Effect of geometrical parameters of venturimeter on pressure drop
.
International Journal of Scientific Research in Science, Engineering and Technology
2
,
865
868
.
Sharp
Z. B.
,
Johnson
M. C.
&
Barfuss
S. L.
2018
Optimizing the ASME Venturi recovery cone angle to minimize head loss
.
Journal of Hydraulic Engineering
144
(
1
),
04017057
.
http://dx.doi.org/10.1061/(ASCE)HY.1943-7900.0001387.
Shi
H.
,
Li
M.
,
Nikrityuk
P.
&
Liu
Q.
2019
Experimental and numerical study of cavitation flows in venturi tubes: from CFD to an empirical model
.
Chemical Engineering Science
207
,
672
687
.
http://dx.doi.org/10.1016/j.ces.2019.07.004.
Shi
H.
,
Li
M.
,
Liu
Q.
&
Nikrityuk
P.
2020
Experimental and numerical study of cavitating particulate flows in a Venturi tube
.
Chemical Engineering Science
219
,
115598
.
https://doi.org/10.1016/j.ces.2020.115598
.
Shook
C. A.
&
Masliyah
J.
1974
Flow of a slurry through a venturi meter
.
The Canadian Journal of Chemical Engineering
52
,
228
233
.
https://doi.org/10.1002/cjce.5450520216.
Tamhakar
N.
,
Pandhare
A.
,
Joglekar
A.
&
Bhansode
V.
2014
Experimental and CFD analysis of flow through venturimeter to determine the coefficient of discharge
.
International Journal of Latest Trends in Engineering and Technology
3
(
4
),
194
200
.
Tukimin
A.
,
Zuber
M.
,
Ahmad
K. A.
2016
CFD analysis of flow through Venturi tube and its discharge coefficient
.
IOP Conference Series: Materials Science and Engineering
152
,
012062
.
https://dx.doi.org/10.1088/1757-899X/152/1/012062
.
von Lavante
E.
,
Zachcial
A.
,
Nath
B.
&
Dietrich
H.
2000
Numerical and experimental investigation of unsteady effects in critical Venturi nozzles
.
Flow Measurement and Instrumentation
11
,
257
264
.
https://doi.org/10.1016/S0955-5986(00)00024-8
.
Weise
J.
,
Baliño
J. L.
&
Paladino
E. E.
2021
CFD study of the transient wet gas flow behavior through orifice plate flow meters
.
Flow Measurement and Instrumentation
82
,
102077
.
https://doi.org/10.1016/j.flowmeasinst.2021.102077
.
Yadav
O.
,
Dahiya
A.
,
Yadav
V. K.
&
Sharma
R.
2022
Experimental and computational investigation of coefficient of discharge of venturimeter
. In:
Recent Trends in Thermal Engineering: Select Proceedings of ICCEMME 2021
(L. M. Das, A. Sharma, F. Y. Hagos & S. Tiwari, eds)
,
Springer
,
Singapore
, pp.
57
72
.
https://doi.org/10.1007/978-981-16-3428-4_6.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).