Next Article in Journal
Technological Advancements for the Detection of Antibiotics in Food Products
Next Article in Special Issue
Demulsification Treatment of Spent Metalworking Fluids by Metal Cations: The Synergistic Effect and Efficiency Evaluation
Previous Article in Journal
Novel One-Step Process for the Production of Bioplastic from Rapeseed Press Cake
Previous Article in Special Issue
Simultaneous Removal of Al, Cu and Zn Ions from Aqueous Solutions Using Ion and Precipitate Flotation Methods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Modification of the Geometry of the Wave-Plate Mist Eliminators on the Droplet Removal Efficiency—CFD Modelling

Faculty of Chemical and Process Engineering, Warsaw University of Technology, Waryńskiego 1, 00-645 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Processes 2021, 9(9), 1499; https://doi.org/10.3390/pr9091499
Submission received: 19 July 2021 / Revised: 23 August 2021 / Accepted: 23 August 2021 / Published: 25 August 2021
(This article belongs to the Special Issue Advanced Liquid Waste and Gas Waste Treatment Processes)

Abstract

:
This study is concerned with droplet separation using wave-plate mist separators. The influence of continuous phase velocity, droplet size, and geometry on droplet removal efficiency has been investigated. The following modifications were analysed: drainage channel presence, length, angle, and modification to the streamlined shape. Furthermore, the influence of the physical parameters of the separated substances on separator efficiency was investigated. Calculations were conducted using computational fluid dynamics (CFD). The results were compared with the experimental data from the literature. Based on the results obtained, a new shape of drainage channels was proposed, which is characterised by high droplet removal efficiency with relatively low pressure drop.

1. Introduction

The necessity of removing liquid droplets from the process gas streams is a common problem in industrial technologies. Rising raw materials prices and strict environmental protection plans have become challenging to design more innovative and efficient systems for the droplet removal from process gases and vapour streams. The presence of gas–liquid aerosols (mist) can result from spontaneous condensation or desublimation in supersaturated gas–vapour mixtures or chemical reactions. It is undesired in many gas-liquid contact devices such as absorbers, quench coolers, or condensers [1]. Eliminating droplets from the gaseous phase aims to prevent the failure of process equipment requiring the use of gases or vapours with high purity, protecting the environment against pollution and degradation, recovering valuable substances, or realising the desire to obtain a product of the highest purity. Wave-plate mist eliminators are applied when necessary to remove droplets of size ranging from a few to tens of micrometres. However, removing very fine droplets requires different solutions such as a glass fibre bed separator intended for removing droplets of a diameter smaller than 2 µm [1]. Wave-plate mist eliminators need to have appropriate geometry (in practice, the channel through which the gas flow has a zig-zag shape, which causes a sharp change in the direction of the flowing fluid stream). The gas containing liquid droplets flows across the bends, forcing it to change the flow direction a few times. Due to their high density compared to gas, liquid droplets are subject to inertial forces, which causes the deflection of the droplet trajectories from the gas motion patch and separation of droplets on the walls. Those forces depend on droplet mass and velocity as well as the geometry of mist eliminator [2]:
F s = m p u p 2 R = π d p 3 ρ p u p 2 6 R
where Fs represents inertial force, R is a bending place curvature radius, and mp, up, dp, and ρp are the mass, velocity, diameter, and density of the droplets, respectively. It can be seen that inertial force is proportional to the d p 3 . Due to this fact, wave-plate mist eliminator efficiencies decrease sharply for smaller droplets. For example, in the same conditions, for 37.5 µm droplets, the separation efficiency of the wave-plate separator was over 90%, but for droplets of diameter less than 25 µm, it dropped sharply to less than 40% [3]. Droplets that have not been re-entrained by the gas can adhere, coalesce, form a film, and flow down under the influence of gravity to be removed from the separator space. The advantage of inertial droplet separators is a low-pressure drop. They can be used to remove foam-forming liquids and high-viscosity liquids, which in the case of mesh and fibre separators is ineffective due to the robust clogging spaces between wires or fibres by liquid droplets. Furthermore, most wave-plate separators have relatively higher resistance to droplet re-entrainment than wire mesh separators, although they cannot obtain high efficiency under conditions of low gas velocities or/and fine droplets [3].
Three essential parameters affect the efficiency of the inertial droplet separator [4]:
  • The linear velocity of the gas;
  • Separator channels’ width;
  • Presence of drainage channels.
As the inertial force value is proportional to the square of the velocity, increasing the gas velocity has a complex influence on droplet removal efficiency. Below 6 m/s, the increase in velocity is beneficial for efficiency; however, for velocities above this value, the influence of secondary droplet entrainment increases, reducing droplet removal efficiency. Nevertheless, for some wave-plate separators with multiple bends, re-entrainment will not occur as long as the gas velocity is lower than 8 m/s [3]. High velocities can also cause flooding phenomenon if a mist eliminator is working in the vertical position. In that case, an upward gas flow prevents the gravitational downflow of droplets [2]. The presence of drainage channels allows maintaining high efficiency even at high linear gas velocities. The shape and type of drainage channels can significantly affect the droplet removal efficiency of the separator.
The problem of optimising the operation of an inertial mist eliminator has been widely described in the literature [5,6,7,8,9,10]; however, an interesting issue is the influence of drainage channel geometry on droplet removal efficiency. Computational fluid dynamics has a unique role in understanding this process. It allows understanding the system’s conditions better and easily estimates the influence of process parameters and geometry changes on droplet removal efficiency.
Some successful CFD analyses of airflow and droplet transport wave-plate mist eliminators are described in the literature [11,12]. There is also proof that the length of drainage channels has a crucial influence on droplet removal efficiency and pressure drop [11]. The novelty of this paper is presenting CFD simulations of the influence of the different aspects of the geometry of drainage pockets such as length, angle, and shape on the droplet removal efficiency and different types of droplets and process medium velocity. The first part of the work is focused on the selection of turbulence and turbulent dispersion models. Next, the crucial part of the work presents the influence of the presence, width, and slope angle of drainage channels. The results concern not only the efficiency but also the pressure drop in the system. Based on the results obtained, a new shape of drainage channels was proposed, which is characterised by a high efficiency of droplet removal with reasonably low pressure drop. In addition, droplet flows of media used in the chemical industry such as n-octane, diesel fuel, sulphuric acid, and aqueous ammonia solution were simulated in the new system.

2. Materials and Methods

2.1. Materials

The present work assumed that air is a continuous phase and water is a discrete phase. Section 3.4 also included parameters of three other media used to simulate discrete phases.

2.2. Experimental System

In the present work, different configurations of commercial wave-plate mist eliminators have been investigated using numerical methods. In the first stage, two types of geometries described by [5] were analysed: one without drainage channels (geometry A) and another with zig-zag shaped drainage channels (geometry B_10_0). Visualisations of considered geometries are presents in Figure 1 and Figure 2. Those results were used to validate chosen numerical methods for further calculations.
In the second stage of work, variants with analogical drainage channels of different lengths (a), which reduce the width of the main channels (L), were simulated. Secondly, there were modelled variants with drainage channels of 10 mm length at different angles (α). Geometrical details of every configuration are listed in Table 1 and presented in Figure 3.

2.3. Simulation Parameters and Assumptions

In order to analyse the effects of using different geometries, medium velocities, and droplet diameters, the commercial software Fluent developed by Ansys Inc. was used. Calculations were carried out using the 2D domain. Calculations were carried out for two different average velocities on the inlet: 2.0 m s−1 and 4.0 m s−1, which are typical values in industrial practice. The upper limit of velocity was determined to avoid dragging droplets.
In order to conduct simulation, the following were assumed:
  • Air is a continuous phase;
  • Water is a discrete phase;
  • Inlet overpressure equal to 5 MPa;
  • Gravitation force on droplets is negligible;
  • No coalescence between droplets;
  • Droplets behave as hard spheres;
  • No droplet-droplet interaction;
  • Droplets do not form a film on the walls;
  • Droplets do not affect velocity profiles of gas phase;
  • Lift forces are negligible;
  • Droplets are removed immediately after hitting a wall;
  • Droplet diameters range from 1 to 40 µm;
  • Mass fraction of water was lower than 0.1.
The design assumptions presented above are commonly used to determine the effect of geometry on system efficiency [3,4,5,11]. In particular, this refers to the value of the mass fraction of water, which below 0.1 should not lead to the formation of a liquid film on the wall, and the phenomenon of droplet coalescence should not significantly affect the system efficiency. The agreement between simulation results and experimental data presented later in this paper confirms the applicability of the model assumptions.
The viscosity and density of continuous phase (air) and discrete phase (water) were specified as constant in constant temperature equal to 15 °C. Those parameters are listed below:
  • ρair = 1.225 kg m−3;
  • νair = 1.461 10−5 m2 s−1;
  • ρwater = 999 kg m−3.

2.4. Continuous Phase Modelling

The SIMPLE method was used for the pressure–velocity coupling, and second-order discretisation schemes were used for all variables to minimise numerical diffusion effects [13]. Computations were regarded as satisfactory converged when total normalised residuals were smaller than 10−6. Computational meshes were generated using Ansys Meshing software and for every variant of geometry. The size of numerical grids was approximately 850,000 quadrilateral cells. The mesh was densest in the wave-plate region, and the near-wall cell sizes were set accordingly to satisfy the y+ ~1 condition. Mesh independence was checked at the highest tested Reynolds numbers using two quantities: average wall shear stress value at the walls and average turbulence energy dissipation rate in the system. The results of both these two quantities were constant (less than 3% difference), even with using denser meshes than those described. The obtained mesh was validated by comparing calculated results with experimental data. The final mesh for the B_10_0 variant was presented in Figure 4.
In the first stage, simulation of B_10_0 geometry was conducted, and obtained results were compared with accurate experimental data with a maximum uncertainty below 5% presented in [5] to select the appropriate turbulence model for further calculations. The gas-phase velocity in a broad part of wave-plate separators is relatively high, and turbulent flow occurs, which means that the k-ε model should be accurate. However, wave-plate channels are narrow, and the flow is less turbulent, making the k-ω model more precise than k-ε in this part of the device. Furthermore, the SST k-ω includes the turbulence shear stress transport, and it is accurate for simulating flow under near-wall conditions [2]. In the present work, three turbulence models based on Reynolds-averaged Navier–Stokes (RANS) were compared:
  • k-ε with standard and enhanced wall treatment;
  • Transition SST k-ω;
  • Reynolds Stress Model (RSM) with enhanced wall treatment.
Many works have been published in recent years presenting the possibility of using more advanced mathematical models, so-called large eddy simulation (LES), to simulate industrial processes [13,14,15,16], enabling the simulation of processes in a wide range of Reynolds numbers. However, it should be remembered that such a solution increases both the calculation time and the demand for computing power many times. It is recommended to use it only when the calculation results are significantly more accurate than those obtained using lower-order models. In a similar process, [17] carried out calculations using RANS and LES models. They observed that differences between models results are 3% and less than 7% on average for the pressure drop and the removal efficiency, respectively, and the RANS approach can be used successfully significantly enhanced with a turbulent dispersion of droplets model.
The following equations govern the continuity and momentum in RANS models:
ρ t + x i ρ u i = 0
t ρ u i + x j ρ u i u j = p x i + x j μ u i x j + u j x i 2 3 δ ij u l x l + x j ρ u i u j ¯ .
For the k-ε model, the turbulent kinetic energy, k, and the specific dissipation rate of k, ε, are obtained from the following equations:
t ρ k + x j ρ ku j = x j μ + μ t σ k k x j + G k + G b ρ ε Y M + S k
t ρ ε + x j ρ ε u j = x j μ + μ t σ ε ε x j + ρ C 1 S ε ρ C 2 ε 2 k + ν ε + C 1 ε ε k C 3 ε G b + S ε .
For the SST k-ω model, the turbulent kinetic energy, k, and the specific dissipation rate of k, ω, are obtained from the following equations:
t ρ k + x i ρ ku i = x j μ + μ t σ k k x j + G k Y k + S k + G b
t ρ ω + x i ρ ω u i = x j μ + μ t σ ω ω x j + G ω Y ω + S ω + G ω b .
The Reynolds Stress Model is more complicated than the other turbulence models mentioned above because of using anisotropic Reynolds stress tensors. It requires most computational resources, and it was used only as a reference.

2.5. Discrete Phase Modelling

Once continuous phase flow is computed, the droplet behaviour can be calculated. In order to simulate droplet motion, a Discrete Phase Model (DPM) was applied. This model uses a Lagrangian approach for modelling droplet trajectories. Droplet motions are tracked using Newton’s second law [2]. The following equation governs forces acting on droplets:
du p dt = F D u u p + g ρ p ρ ρ p + F x
where u and up are gas and droplet velocities, respectively; ρ and ρp are gas and droplet densities, respectively; g is gravitational acceleration, and Fx is an additional acceleration. In this work, one-way coupling between continuous and discrete phases was considered. It is assumed that droplet movement does not affect the continuous phase flow pattern. In Equation (8), FD(u − up) is a resistant force of continuous phase per droplet mass (m s−2), and it can be calculated from the Reynolds number and droplet parameters. Coefficient FD can be expressed as:
F D = 18 µ ρ p d p 2 C D Re 24
where Re is the relative Reynolds number, and CD is the drag force coefficient, which is obtained from the following expressions [18]:
C D = 0.44   for   Re d > 1000
C D = 1 + 0.15 Re d 0.687 Re d 24   for   Re d 1000
where Red is the droplet Reynolds number defined by:
Re d = ρ u u p d p µ .
For efficiency calculations, we used droplets of size within the range from 1 µm to the size when the calculated efficiency equals 100%. In all cases, droplets were injected in a perpendicular direction to the inlet plane. Efficiency is defined as a ratio of the number of droplets removed from continuous phase (cin − cout) to the total injected droplets (cin):
η = c in c out c in   · 100 % .
The droplet motion was calculated using two variants of the DPM model:
  • Without turbulent dispersion;
  • With turbulent dispersion affecting droplet motion.
To include the phenomenon of turbulent dispersion in simulating droplet motion, stochastic tracking model was used, which allows calculating droplet motion trajectories by using instantaneous velocity:
u t =   u ¯ + u t .
Many particles are injected at one point in a stochastic model, and their motion trajectories are calculated using a fluctuation module.
The results of all RANS models described in Section 2.4 were verified using experimental data [5]. Figure 5 and Table 2, Table 3, Table 4 and Table 5 show a summary of results and relative errors for geometry B_10_0, which has the same structure and size as used in the work of [5]. In this case, the turbulent dispersion model was implemented to improve turbulent models used for calculations. The influence of including this model in calculations will be presented in the next section.
Analysing Figure 5 and Table 2, Table 3, Table 4 and Table 5, one can assume that all used models predict the trend of increase in efficiency with droplet size. However, there are significant differences in expected values. The most accurate results were observed using the SST model, especially for higher efficiencies. Furthermore, the SST model is the only one that accurately predicts almost 100% efficiency for droplets with a diameter of 18 µm (2 m s−1) and 12 µm (4 m s−1), which are very close results to experimental data. This conclusion is similar to [5] and [15]. As a result, the SST model was found as the most suitable for simulating the gas phase.

2.6. Comparing Calculations with and without Turbulent Dispersion

In Figure 6 and Table 6, Table 7, Table 8 and Table 9, the results and relative errors with and without using turbulent dispersion for the SST model are presented and compared with experimental data.
Analysing Figure 6 and Table 6, Table 7, Table 8 and Table 9, one can observe that more accurate results, compared to the experimental results, were received using the SST model with turbulent dispersion, particularly for droplets of small diameter obtaining higher efficiency, which is following expectations, because for smaller droplets, the model predicts a higher acceleration of droplets, impacting fluctuations of the continuous phase. Based on the obtained results, all subsequent calculations will be done using the SST model with turbulent dispersion as the most suitable model describing the work of wave-plate mist eliminator.

2.7. Non-Dimensional Numbers

Calculated values of Reynolds number on the inlet, for hydraulic diameter dh = 0.68 m, are presented in Table 10.
The Reynolds number is the most important non-dimensional number considered case, because it illustrates the nature of gas flow (Re) and droplet flow (Red). It allows one to select the appropriate turbulence model for CFD calculations and parameters used in the equations. However, the Stokes number, representing the ratio between droplet relaxation time and gas characteristic time, is also loosely related to droplet removal efficiency, even though it is not related to turbulent dispersion effects [5]. Since the characteristic length is channel width, the Stokes number can be calculated by the following equation:
St = ρ p d p 2 u in sin α c 18 µ L   sin   α c
where uin is gas velocity on inlet and αc is the bend angle between wave-plates. The relation between Stokes number and removal efficiency is presented in Figure 7. It can be noticed that the regression curve shape is similar to the droplet diameter–efficiency curve obtained from calculations.

2.8. Boundary Conditions

For the continuous phase, the following boundary conditions were used:
  • “Velocity-inlet” with constant value on the inlet to the computational domain;
  • “Outflow” on the outlet of the computational domain;
  • “Stationary wall” without slip on the edges of the wave-plate mist eliminator and drainage channels;
  • “Symmetry” on the edges of the computational domain outside the wave-plate mist eliminator excluding inlet and outlet.
For the discrete phase, the following boundary conditions were used:
  • “Trap” on the edges of the wave-plate mist eliminator and drainage channels;
  • “Escape” on the rest of the edges of the computational domain.
Droplets of the dispersed phase were put into the domain through the inlet by the uniform injections of spheres separately for every diameter from investigated range (1 to 40 µm, every 3 µm) with corresponding velocity and constant total mass flow rate of 0.00355 kg s−1. Every injection was calculated separately.

3. Results

From a practical and industrial point of view, three parameters were crucial to simulate flow fields of analysed geometries:
  • The removal efficiency of droplets;
  • Pressure drop on wave-plate mist-eliminator;
  • The mean force acting on the single drainage channel.
Pressure drop was calculated as a difference between pressures on the inlet and outlet of investigated mist eliminators. The mean force acting on the drainage channel was calculated as an arithmetic average from forces acting on every drainage channel. Calculation results for both parameters are listed in Table 11 and Table 12.
Based on the above results, it can be assumed that according to Darcy–Weisbach law, the pressure drop (Δp) increases proportionally to the square of the medium velocity (u):
Δ p ~ u 2 .
The results of pressure drop are following the literature data [2,14]. From an economic point of view, the pressure drop is an undesirable phenomenon. As a result, geometry variant B_30_0 most likely has no practical usage. Furthermore, it can be observed that the force acting on the drainage channel is proportional to the pressure drop on the mist eliminator. In the following chapters, the influence of various parameters was considered in detail.

3.1. Influence of Gas Inlet Velocity on Droplet Removal Efficiency on Wave-Plate Mist Eliminator

The calculations were carried out for two inlet velocities, 2 m s−1 and 4 m s−1, and two characteristic geometries were investigated:
  • A;
  • B_10_0.
The obtained results are shown in Figure 8.
The results confirm that the increase of the inlet velocity of the gas increases the droplet separation efficiency. In B_10_0, it can be seen that for the speed of 4 m s−1, almost 100% efficiency is obtained for the droplet diameter of 12 µm (18 µm for the lower velocity). Nevertheless, it should be noted that this is done at the cost of more than four times higher pressure drop. The results also confirm that drainage channels positively affect separation efficiency for larger droplet diameters, increasing it by about 20–30%. For droplet sizes smaller than 9 µm for 2 m s−1 and 7 µm for 4 m s−1, the drop separation efficiency is higher for a system without channels, which is obviously due to the inertia forces that act on the drops. In a system with drainage channels, the cross-section near the channels decreases, which causes an increasing velocity, and smaller drops follow the gas rather than deposition on the walls.

3.2. Influence of Drainage Channels Length on Droplet Removal Efficiency on Wave-Plate Mist Eliminator

The next step was to investigate the influence of the length of drainage channels crucial for the mist eliminator. In the present work, we investigated geometry variants with drainage channels with the following lengths:
  • 10.5 mm—geometry B_10_0;
  • 15.5 mm—geometry B_15_0;
  • 20.5 mm—geometry B_20_0;
  • 25.5 mm—geometry B_25_0;
  • 30.5 mm—geometry B_30_0.
The obtained results are shown in Figure 9.
Based on the results, the droplet removal efficiency increases with drainage channel length. This phenomenon is associated with the increased velocity of the continuous phase, decreased area of flow, and increased area of the drainage channel. Figure 10 and Figure 11 show velocity fields and vectors of investigated geometries. It can be observed that the highest gas velocities are obtained in the close neighbourhood to drainage channels and that the velocity increases with the next channel, which means that self-similarity flow conditions for the subsequent cross-sections near the channel have not been achieved. These results are similar to those of [5].
With areas with high-velocity values, stagnation areas and recirculation areas are observed in and just behind drainage channels. All observed phenomena cause the drainage channels with higher lengths to provide better results, although it also generates higher pressure drops, which are undesirable. Those variants of geometries can be used if high-pressure drops are acceptable.

3.3. Influence of Drainage Channels Angle on Droplet Removal Efficiency on Wave-Plate Mist Eliminator

Simulations were conducted for geometry variants with different angles of drainage channels. Geometry B_10_0 was modified to obtain three different angles of drainage channels:
  • 30°—geometry B_10_30;
  • 60°—geometry B_10_60;
  • 90°—geometry B_10_90.
The results in Figure 12, Figure 13 and Figure 14 show the velocity fields of the investigated geometries.
Droplet removal efficiency increases with drainage channel angle (α). This phenomenon can result from moving drainage channels to areas with higher gas velocities, which causes higher inertial force acting on droplets in those areas and generates higher local pressure drops. The maximum velocity increases with the slope of the channels. In B_10_0, the maximum velocity is 5.5 times higher than the system inlet velocity, and in B_10_90, it is 8.5 times higher. Simultaneously, the separation efficiency for drainage channel slopes of 0 and 30 degree increases by about 40%, with the pressure drop by about 25%. Higher angles can also increase secondary droplet entrainment; however, the model does not include this effect.

3.4. Influence of Physicochemical Properties of the Droplet on Removal Efficiency on Wave-Plate Mist Eliminator

Mist eliminators are used for removing droplets of media with different physicochemical properties. In the present work, we analysed the influence of different types of droplet materials on the work of the mist eliminator listed in Table 13.
Calculations were carried out for geometry variants A and B_10_0. The results are shown in Figure 15.
Based on the plots shown in Figure 15, it can be observed that the density of the medium has a strong influence on the droplet removal efficiency. This phenomenon is since the inertia force acting on droplets is directly proportional to droplet mass. A higher density of medium causes better droplet removal efficiency. A similar effect can be observed by analysing the change in droplet size. Higher diameters of droplets cause higher droplet mass and hence inertia force, making them easier to remove by a mist eliminator. The presented results again show a better droplet removal efficiency when using a system with drainage channels, where the efficiency increases by about 20%. However, it should be remembered that this happens at the cost of an increased pressure drop of about 6%. A more visible effect is observed for larger droplet sizes.

3.5. Modification of Wave-Plate Mist Eliminator with Streamline Drainage Channels

The results presented in the previous sections show that the computational fluid dynamics correctly predict the influence of process and geometric parameters on the removal efficiency. After analysing the results, the authors proposed a new drainage channel to achieve high process efficiency with a slight pressure drop. When analysing the pressure distribution for the cases under consideration, e.g., geometry B_15_0 shown in Figure 16, it can be seen that high-pressure drop values appear near the sharp edges of the drainage walls, predominantly perpendicular to the flow direction.
For longer edges, this effect is even more visible. Therefore, it was decided to propose drainage channels with a much more streamlined shape. The geometry of the new system is shown in Figure 17, and the main dimensions are shown in Table 14.
The results of droplet removal efficiency are shown in Figure 18 and Figure 19. The authors decided to compare the results obtained for the new geometry for the most similar previous cases: B_10_0 and B_15_0.
Based on the plots presented in Figure 16, geometry variant C_15_0 provides the highest droplet removal efficiency.
Figure 19 shows velocity fields and vectors of the investigated new geometries, and Table 15 and Table 16 present the pressure drop and the mean force acting on drainage channels for the three, comparing geometry. For comparison, Figure 20 shows velocity fields and vectors for the analogous case with non-streamline drainage channels.
Analysing the velocity distributions of the considered cases for 2 m s−1, one can observe that the new geometry gets higher maximum velocities than the corresponding variant with non-streamlined drainage channels. This phenomenon is related to the shape of the external drainage channel walls. In C_15_0, the main gas stream is flattened, which is mainly visible on the third channel flows. A critical phenomenon is the occurrence of vortexes inside the channels. In the case of simple channels, the vortex is flattened in an oval shape. Around the channel’s edges, the fluid from inside comes into contact with a gas with a nearly maximum velocity. Although, in the case of the profiled channels, a different phenomenon occurs. The forming vortex is round in shape, and the channel’s edge disturbs the flow to a lesser extent, which will substantially impact the entrapment of deposited droplets—this can be well observed analysing the droplet trajectories presented in Figure 21 and Figure 22.
Based on the visible results, the droplet removal efficiency is about 100% for droplets in the size of 14 µm. This new variant also has over two times higher droplet removal efficiency than geometry variant B_10_0 for droplets of 6 µm and 80% higher droplet removal than variant B_15_0. However, the pressure drop is about 23% higher than variant B_10_0 and similar to variant B_15_0 for the inlet velocity of 2 m s−1. One can observe that the results are even better for higher velocity. The streamline construction of drainage channels should also minimalise the phenomenon of secondary droplet entrainment. Naturally, experiments should validate every result obtained through simulation.

4. Conclusions

The work of the inertial droplet separator was modelled using computational fluid mechanics. The influence of practical important parameters on the droplet removal efficiency was investigated: the inlet velocity of the continuous phase, the droplet size, the effect of the presence of drainage channels, and their geometry, i.e., the length and angle of the slope. The simulation results show that drainage channels increase the pressure drop, which is undoubtedly a disadvantage of using drainage channels in inertial droplet separators. On the other hand, the removal efficiency values of the separator with drainage channels are much higher than those without drainage channels. Drainage channels cause the velocity of droplets to increase, which results in increased removal efficiency. At the same time, it should be remembered that the literature data show that the effect of droplet entrainment (not included in the calculation) will be much higher in the separator without drainage channels, which will further reduce the separation efficiency.
Three turbulence models and two variants of turbulent dispersion modelling were used in the present work. The best results were obtained for the transition SST model with the turbulent dispersion model attached. Such conclusions are consistent with the literature, and an excellent agreement with the experimental literature data allows simulating other parameters such as changes in slope angle and the size of drainage channels.
The simulation results show that small changes in geometry significantly improve the separator efficiency with a relatively small increase in pressure drop. The analysis of the obtained results allowed authors to propose a new geometry of drainage channels with streamlined shapes. This solution allows achieving a significant increase in droplet removal efficiency while keeping a comparable pressure drop for lower velocities and even decreasing for higher velocities than the basic geometry of channels of similar length. These results should be confirmed by extending the theoretical model, particularly with the effects of secondary entrapment of droplets and forming a liquid film on the wall, and comparing these with the result of the experimental data.

Author Contributions

Conceptualisation, Ł.M. and J.Ł.; methodology, Ł.M.; validation, Ł.M. and J.Ł.; formal analysis, Ł.M.; investigation Ł.M., J.Ł., M.T.; writing—original draft preparation, M.T.; writing—review and editing, Ł.M. and M.T.; visualisation, M.T.; supervision, Ł.M.; project administration, Ł.M. 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

Data available on request.

Conflicts of Interest

The authors declare no conflict of interest.

Symbols

The following symbols are used in this manuscript:
Latin symbols:
adrainage channel length, mm
coutnumber of droplets removed from the system
cinnumber of droplets injected into the system
CDdrag force coefficient
dpdroplet diameter, m
dhhydraulic diameter of the inlet
FDcoefficient in drag force acceleration, s−1
Fsdroplet inertial force, N
ggravitational acceleration, m s−2
kturbulent kinetic energy, m2 s−2
Lmain channel width, mm
mpdroplet mass, kg
ppressure, Pa
Rcurvature radius of the bending place, m
ReReynolds number
Reddroplet Reynolds number
StStokes number
ttime, s
ucontinuous phase velocity, m s−1
u ¯ continuous phase average velocity, m s−1
u’fluctuation of continuous phase velocity, m s−1
uincontinuous phase velocity on inlet, m s−1
updroplet velocity, m s−1
y+dimensionless wall distance
Greek symbols:
αdrainage channel angle, °
αcangle between wave-plates, °
δdelta function
εturbulent dissipation rate, m2 s−3
ηdroplet removal efficiency
µcontinuous phase dynamic viscosity, Pa·s
μtcontinuous phase turbulent viscosity, Pa·s
νaircontinuous phase kinematic viscosity, m2 s−1
ρ; ρaircontinuous phase density, kg m−3
ρp; ρwaterdroplet density, kg m−3
σkturbulent Prandtl number for k
σεturbulent Prandtl number for ε
σkturbulent Prandtl number for ω
ωspecific dissipation rate, s−1
Acronyms
CFDComputational Fluid Dynamics
DPMDisperse Phase Model
LESLarge Eddy Simulation
RANSReynolds-Averaged Navier–Stokes
RSM, RSReynolds Stress Model
SIMPLESemi-Implicit Method for Pressure Linked Equations
SSTShear Stress Transport

References

  1. Del Ministro, C.; Bosio, F.; Spreafico, S.; Brunazzi, E. The gx fiber bed—A novel mist eliminator with improved aerosol removal performances. Chem. Eng. Trans. 2018, 69, 811–816. [Google Scholar] [CrossRef]
  2. Yuan, S.; Fan, Y.; Li, J.; Zhou, S.; Cao, Y. Influence of droplet coalescence and breakup on the separation process in wave-plate separators. Can. J. Chem. Eng. 2018, 96, 1627–1636. [Google Scholar] [CrossRef]
  3. Liu, Y.; Yu, D.; Jiang, J.; Yu, X.; Yao, H.; Xu, M. Experimental and numerical evaluation of the performance of a novel compound demister. Desalination 2017, 409, 115–127. [Google Scholar] [CrossRef]
  4. Gharib, J.; Moraveji, M.K. Determination the factors affecting the vane-plate demisters efficiency using CFD modeling. J. Chem. Eng. Process Technol. 2011, 1, 1–5. [Google Scholar] [CrossRef] [Green Version]
  5. Galletti, C.; Brunazzi, E.; Tognotti, L. A numerical model for gas flow and droplet motion in wave-plate mist eliminators with drainage channels. Chem. Eng. Sci. 2008, 63, 5639–5652. [Google Scholar] [CrossRef]
  6. Azzopardi, B.J.; Sanaullah, K.S. Re-entrainment in wave-plate mist eliminators. Chem. Eng. Sci. 2002, 57, 3557–3563. [Google Scholar] [CrossRef]
  7. Houghton, H.; Radford, W. Measurement on eliminators and the development of a new type for use at high gas velocities. Trans. Am. Inst. Chem. Eng. 1939, 35, 427–433. [Google Scholar]
  8. James, P.; Wang, Y.; Azzopardi, B.; Hughes, J. The role of drainage channels in the performance of wave-plate mist eliminators. Chem. Eng. Res. Des. 2003, 81, 639–648. [Google Scholar] [CrossRef]
  9. Wang, Y.; James, P. The calculation of wave-plate demister efficiencies using numerical simulation of the flow field and droplet motion. Chem. Eng. Res. Des. 1998, 76, 980–985. [Google Scholar] [CrossRef]
  10. Wang, Y.; James, P. Assessment of an eddy-interaction model and its refinements using predictions of droplet deposition in a wave-plate demister. Chem. Eng. Res. Des. 1999, 77, 692–698. [Google Scholar] [CrossRef]
  11. Estakhrsar, M.H.H.; Rafee, R. Effect of drainage channel dimensions on the performance of wave-plate mist eliminators. Korean J. Chem. Eng. 2013, 30, 1301–1311. [Google Scholar] [CrossRef]
  12. Rafee, R.; Rahimzadeh, H.; Ahmadi, G. Numerical simulations of airflow and droplet transport in a wave-plate mist eliminator. Chem. Eng. Res. Des. 2010, 88, 1393–1404. [Google Scholar] [CrossRef]
  13. Michalcová, V.; Kotrasová, K. The numerical diffusion effect on the CFD simulation accuracy of velocity and temperature field for the application of sustainable architecture methodology. Sustainability 2020, 12, 10173. [Google Scholar] [CrossRef]
  14. Makowski, Ł.; Bałdyga, J. Large eddy simulation of mixing effects on the course of parallel chemical reactions and comparison with k–ε modeling. Chem. Eng. Process. Process Intensif. 2011, 50, 1035–1040. [Google Scholar] [CrossRef]
  15. Makowski, Ł.; Orciuch, W.; Bałdyga, J. Large eddy simulations of mixing effects on the course of precipitation process. Chem. Eng. Sci. 2012, 77, 85–94. [Google Scholar] [CrossRef]
  16. Wojtas, K.; Makowski, Ł.; Orciuch, W. Barium sulfate precipitation in jet reactors: Large eddy simulations, kinetics study and design considerations. Chem. Eng. Res. Des. 2020, 158, 64–76. [Google Scholar] [CrossRef]
  17. Ruiz, J.; Kaiser, A.; Zamora, B.; Cutillas, C.; Lucas, M. CFD analysis of drift eliminators using RANS and LES turbulent models. Appl. Therm. Eng. 2016, 105, 979–987. [Google Scholar] [CrossRef]
  18. Gosman, A.D.; Loannides, E. Aspects of computer simulation of liquid-fueled combustors. J. Energy 1983, 7, 482–490. [Google Scholar] [CrossRef]
Figure 1. Wave-plate mist eliminator with inflow and outflow area (geometry A) dimensions in mm.
Figure 1. Wave-plate mist eliminator with inflow and outflow area (geometry A) dimensions in mm.
Processes 09 01499 g001
Figure 2. Visualisation of wave-plate mist eliminator with drainage channels (geometry B_10_0).
Figure 2. Visualisation of wave-plate mist eliminator with drainage channels (geometry B_10_0).
Processes 09 01499 g002
Figure 3. Details of geometries of analysed wave-plate mist eliminators dimensions in mm.
Figure 3. Details of geometries of analysed wave-plate mist eliminators dimensions in mm.
Processes 09 01499 g003
Figure 4. Final computational mesh of geometry B_10_0: (a) view on a fragment of mist eliminator; (b) detailed view on drainage channel.
Figure 4. Final computational mesh of geometry B_10_0: (a) view on a fragment of mist eliminator; (b) detailed view on drainage channel.
Processes 09 01499 g004aProcesses 09 01499 g004b
Figure 5. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 5. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g005
Figure 6. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 6. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g006
Figure 7. Comparison of calculated Stokes numbers with results obtained with the SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 7. Comparison of calculated Stokes numbers with results obtained with the SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g007
Figure 8. Comparison of calculated results of droplet removal efficiency for geometries A and B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 8. Comparison of calculated results of droplet removal efficiency for geometries A and B_10_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g008
Figure 9. Comparison of calculated results of droplet removal efficiency for geometries with different drainage channel lengths with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 9. Comparison of calculated results of droplet removal efficiency for geometries with different drainage channel lengths with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g009
Figure 10. Velocity fields (a,c) (m s−1) and vectors (b,d) and vectors (m s−1) of investigated geometry variants with different drainage channel lengths: (a,b) B_10_0; (c,d) B_25_0; with inlet velocity 2 m s−1 simulated using a transition SST model.
Figure 10. Velocity fields (a,c) (m s−1) and vectors (b,d) and vectors (m s−1) of investigated geometry variants with different drainage channel lengths: (a,b) B_10_0; (c,d) B_25_0; with inlet velocity 2 m s−1 simulated using a transition SST model.
Processes 09 01499 g010aProcesses 09 01499 g010b
Figure 11. Velocity fields (a,c) (m s−1) and vectors (b,d) and vectors (m s−1) of investigated geometry variants with different drainage channel lengths: (a,b) B_25_0; (c,d) B_30_0; with inlet velocity 4 m s−1 simulated using a transition SST model.
Figure 11. Velocity fields (a,c) (m s−1) and vectors (b,d) and vectors (m s−1) of investigated geometry variants with different drainage channel lengths: (a,b) B_25_0; (c,d) B_30_0; with inlet velocity 4 m s−1 simulated using a transition SST model.
Processes 09 01499 g011aProcesses 09 01499 g011b
Figure 12. Comparison of calculated results of droplet removal efficiency for geometries with different drainage channel angles with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 12. Comparison of calculated results of droplet removal efficiency for geometries with different drainage channel angles with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g012
Figure 13. Velocity fields (m s−1) of investigated geometry variants with different drainage channel angles: (a) B_10_30; (b) B_10_60; (c) B_10_90; for inlet velocity 2 m s−1, simulated using a transition SST model.
Figure 13. Velocity fields (m s−1) of investigated geometry variants with different drainage channel angles: (a) B_10_30; (b) B_10_60; (c) B_10_90; for inlet velocity 2 m s−1, simulated using a transition SST model.
Processes 09 01499 g013aProcesses 09 01499 g013b
Figure 14. Velocity fields (m s−1) of investigated geometry variants with different drainage channel angles: (a) B_10_30; (b) B_10_60; (c) B_10_90; for inlet velocity 4 m s−1, simulated using a transition SST model.
Figure 14. Velocity fields (m s−1) of investigated geometry variants with different drainage channel angles: (a) B_10_30; (b) B_10_60; (c) B_10_90; for inlet velocity 4 m s−1, simulated using a transition SST model.
Processes 09 01499 g014
Figure 15. Comparison of calculated results of droplet removal efficiency for different media: (a) geometry A, inlet velocity 2 m s−1; (b) geometry A, inlet velocity 4 m s−1; (c) geometry B_10_0, inlet velocity 2 m s−1; (d) geometry B_10_0, inlet velocity 4 m s−1.
Figure 15. Comparison of calculated results of droplet removal efficiency for different media: (a) geometry A, inlet velocity 2 m s−1; (b) geometry A, inlet velocity 4 m s−1; (c) geometry B_10_0, inlet velocity 2 m s−1; (d) geometry B_10_0, inlet velocity 4 m s−1.
Processes 09 01499 g015aProcesses 09 01499 g015b
Figure 16. Pressure fields (Pa) of B_15_0 geometry for inlet velocity 2 m s−1 simulated using a transition SST model.
Figure 16. Pressure fields (Pa) of B_15_0 geometry for inlet velocity 2 m s−1 simulated using a transition SST model.
Processes 09 01499 g016
Figure 17. Details of geometry C_15_0 with streamlined drainage channels.
Figure 17. Details of geometry C_15_0 with streamlined drainage channels.
Processes 09 01499 g017
Figure 18. Comparison of calculated results of droplet removal efficiency for geometries B_10_0, B_15_0, and C_15_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Figure 18. Comparison of calculated results of droplet removal efficiency for geometries B_10_0, B_15_0, and C_15_0 with inlet velocity: (a) 2 m s−1; (b) 4 m s−1.
Processes 09 01499 g018
Figure 19. Velocity fields (m s−1) of geometry variant C_15_0 with streamline drainage channels with inlet velocities (a,b) 2 m s−1 and (c,d) 4 m s−1, which were simulated using a transition SST model. Velocity fields (a,c) (m s−1) and vectors (b,d) (m s−1) of geometry variant C_15_0 with streamline drainage channels with inlet velocities (a,b) 2 m s−1 and (c,d) 4 m s−1, which were simulated using a transition SST model.
Figure 19. Velocity fields (m s−1) of geometry variant C_15_0 with streamline drainage channels with inlet velocities (a,b) 2 m s−1 and (c,d) 4 m s−1, which were simulated using a transition SST model. Velocity fields (a,c) (m s−1) and vectors (b,d) (m s−1) of geometry variant C_15_0 with streamline drainage channels with inlet velocities (a,b) 2 m s−1 and (c,d) 4 m s−1, which were simulated using a transition SST model.
Processes 09 01499 g019aProcesses 09 01499 g019bProcesses 09 01499 g019c
Figure 20. Velocity fields (a) and vectors (b) (m s−1) (a) and vectors (b) (m s−1) of geometry variant B_15_0 with inlet velocities 2 m s−1 simulated using a transition SST model.
Figure 20. Velocity fields (a) and vectors (b) (m s−1) (a) and vectors (b) (m s−1) of geometry variant B_15_0 with inlet velocities 2 m s−1 simulated using a transition SST model.
Processes 09 01499 g020aProcesses 09 01499 g020b
Figure 21. Droplet trajectories for geometries B_15_0 (a,c,e,g) and C_15_0 (b,d,f,h) for velocity 2 (m s−1) and droplet diameters: 8 (µm) (a,b); 10 (µm) (c,d); 12 (µm) (e,f); 14 (µm) (g,h).
Figure 21. Droplet trajectories for geometries B_15_0 (a,c,e,g) and C_15_0 (b,d,f,h) for velocity 2 (m s−1) and droplet diameters: 8 (µm) (a,b); 10 (µm) (c,d); 12 (µm) (e,f); 14 (µm) (g,h).
Processes 09 01499 g021
Figure 22. Droplet trajectories for geometries B_15_0 (a,c,e,g) and C_15_0 (b,d,f,h) for velocity 4 (m s−1) and droplet diameters: 6 (µm) (a,b); 8 (µm) (c,d); 10 (µm) (e,f); 12 (µm) (g,h).
Figure 22. Droplet trajectories for geometries B_15_0 (a,c,e,g) and C_15_0 (b,d,f,h) for velocity 4 (m s−1) and droplet diameters: 6 (µm) (a,b); 8 (µm) (c,d); 10 (µm) (e,f); 12 (µm) (g,h).
Processes 09 01499 g022
Table 1. Geometrical details of analysed mist-wave eliminators.
Table 1. Geometrical details of analysed mist-wave eliminators.
GeometryPresence of Drainage Channelsa [mm]L [mm]α [°]
ANo-35.5-
B_10_0Yes10.5250
B_10_30Yes10.52530
B_10_60Yes10.52560
B_10_90Yes10.52590
B_15_0Yes15.5200
B_20_0Yes20.5150
B_25_0Yes25.5100
B_30_0Yes30.550
Table 2. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Table 2. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Droplet Diameter [µm]Experimental Data [5] [-]k-ε, std. Wall tr. [-]k-ε, enh. Wall tr. [-]RS, enh. Wall tr. [-]SST with turb. disp. [-]
80.200.7660.4870.6050.270
100.480.8070.6190.7170.468
120.750.8090.6880.7620.695
130.930.8190.7320.8020.798
150.980.8560.8280.8950.946
180.990.9010.9130.9540.996
2110.9440.9680.9921
Table 3. Relative errors between calculated results and experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Table 3. Relative errors between calculated results and experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Droplet Diameter [µm]k-ε, std. Wall tr. [%]k-ε, enh. Wall tr. [%]RS, enh. Wall tr. [%]SST with turb. disp. [%]
8282.98143.40202.2934.91
1068.2128.8649.412.58
127.878.271.67.33
1311.8821.2713.7614.18
1512.6515.518.673.47
188.997.793.640.61
215.603.200.800
Table 4. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Table 4. Comparison of calculated results with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Droplet Diameter [µm]Experimental Data [5] [-]k-ε, std. Wall tr. [-]k-ε, enh. Wall tr. [-]RS, enh. Wall tr. [-]SST with turb. disp. [-]
60.200.7590.6160.6340.328
70.400.7970.7070.7190.493
80.680.8370.7890.7990.686
100.860.8830.8830.9020.944
110.950.8870.8970.9240.982
1310.9040.9220.9590.995
Table 5. Relative errors between calculated results and experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Table 5. Relative errors between calculated results and experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Droplet Diameter [µm]k-ε, std. Wall tr. [%]k-ε, enh. Wall tr. [%]RS, enh. Wall tr. [%]SST with turb. disp. [%]
6279.50208.00217.0064.00
799.1376.6579.8023.37
823.1416.0717.470.88
102.682.734.839.76
116.665.602.703.34
139.577.774.060.5
Table 6. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Table 6. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Droplet Diameter [µm]Experimental Data [5] [-]SST without turb. disp. [-]SST with turb. disp. [-]
80.200.0720.270
100.480.1380.468
120.750.2280.695
130.930.4850.798
150.9810.946
180.9910.996
21111
Table 7. Relative errors between calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Table 7. Relative errors between calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 2 m s−1.
Droplet Diameter [µm]SST without turb. disp. [%]SST with turb. disp. [%]
863.8334.91
1071.252.58
1269.607.33
1347.8114.18
152.043.47
181.010.61
2100
Table 8. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Table 8. Comparison of calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Droplet Diameter [µm]Experimental Data [5] [-]SST without turb. disp. [-]SST with turb. disp. [-]
60.200.0800.328
70.400.2760.493
80.680.4710.686
100.860.7780.944
110.950.8890.982
13110.995
Table 9. Relative errors between calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Table 9. Relative errors between calculated results obtained with SST model with and without using turbulent dispersion with experimental data for geometry B_10_0 with inlet velocity 4 m s−1.
Droplet Diameter [µm]SST without turb. disp. [%]SST with turb. disp. [%]
660.0064.00
731.0823.37
830.690.88
109.549.76
116.423.34
1300.5
Table 10. Reynolds number values for different variants of calculation.
Table 10. Reynolds number values for different variants of calculation.
The Average Velocity on the Inlet [m s−1]Reynolds Number [-]
2.094,114
4.0188,277
Table 11. Pressure drops on mist eliminators and the mean force acting on drainage channel; inlet velocity 2 m s−1.
Table 11. Pressure drops on mist eliminators and the mean force acting on drainage channel; inlet velocity 2 m s−1.
GeometryAB_10_0B_10_30B_10_60B_10_90B_15_0B_20_0B_25_0B_30_0
Pressure drop [Pa]82871061512051092465871977
Mean force acting on drainage channel [N]-0.380.480.630.781.384.2011.6443.61
Table 12. Pressure drops on mist eliminators and the mean force acting on drainage channel; inlet velocity 4 m s−1.
Table 12. Pressure drops on mist eliminators and the mean force acting on drainage channel; inlet velocity 4 m s−1.
GeometryAB_10_0B_10_30B_10_60B_10_90B_15_0B_20_0B_25_0B_30_0
Pressure drop [Pa]346367446643872442100023717960
Mean force acting on drainage channel [N]-1.531.962.593.265.6217.1147.07175.81
Table 13. Materials of droplets used in calculations and their densities.
Table 13. Materials of droplets used in calculations and their densities.
MediumDensity [kg m−3]
water999
n-octane720
diesel fuel830
95% sulphuric acid1834
25% aqueous ammonia solution865
Table 14. Geometrical details of analysed mist-wave eliminator with streamlined drainage channels.
Table 14. Geometrical details of analysed mist-wave eliminator with streamlined drainage channels.
NamePresence of Drainage Channelsa [mm]L [mm]
C_15_0Yes15.518.5
Table 15. Pressure drops on mist eliminators and the mean force acting on the drainage channel; inlet velocity 2 m s−1.
Table 15. Pressure drops on mist eliminators and the mean force acting on the drainage channel; inlet velocity 2 m s−1.
GeometryB_10_0B_15_0C_15_0
Pressure drop [Pa]87109107
Mean force acting on drainage channel [N]0.381.381.40
Table 16. Pressure drops on mist eliminators and the mean force acting on drainage the channel; inlet velocity 4 m s−1.
Table 16. Pressure drops on mist eliminators and the mean force acting on drainage the channel; inlet velocity 4 m s−1.
GeometryB_10_0B_15_0C_15_0
Pressure drop [Pa]367442406
A mean force acting on drainage channel [N]1.535.626.20
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Makowski, Ł.; Łaskowski, J.; Tyrański, M. Influence of Modification of the Geometry of the Wave-Plate Mist Eliminators on the Droplet Removal Efficiency—CFD Modelling. Processes 2021, 9, 1499. https://doi.org/10.3390/pr9091499

AMA Style

Makowski Ł, Łaskowski J, Tyrański M. Influence of Modification of the Geometry of the Wave-Plate Mist Eliminators on the Droplet Removal Efficiency—CFD Modelling. Processes. 2021; 9(9):1499. https://doi.org/10.3390/pr9091499

Chicago/Turabian Style

Makowski, Łukasz, Jakub Łaskowski, and Mariusz Tyrański. 2021. "Influence of Modification of the Geometry of the Wave-Plate Mist Eliminators on the Droplet Removal Efficiency—CFD Modelling" Processes 9, no. 9: 1499. https://doi.org/10.3390/pr9091499

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