Next Article in Journal
Design of Periodic Neural Networks for Computational Investigations of Nonlinear Hepatitis C Virus Model Under Boozing
Next Article in Special Issue
Non-Hydrostatic Galerkin Model with Weighted Average Pressure Profile
Previous Article in Journal
A DNN-Based Surrogate Constitutive Equation for Geometrically Exact Thin-Walled Rod Members
Previous Article in Special Issue
Transport Characteristics of Small Molecules Diffusing near Deforming Blood Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Computational Analysis of Pipe Roughness Influence on Slurry Flow Dynamics

1
Department of Mechanical and Aerospace Engineering, Monash University Clayton, Melbourne, VIC 3800, Australia
2
Government Industrial Training Institute, Chautala, Sirsa 125101, India
3
Department of Mechanical Engineering, University of the Philippines Los Baños, Los Baños 4031, Philippines
4
School of Innovation, Design and Technology, Wellington Institute of Technology, Wellington 5012, New Zealand
*
Author to whom correspondence should be addressed.
Computation 2025, 13(3), 65; https://doi.org/10.3390/computation13030065
Submission received: 12 November 2024 / Revised: 10 February 2025 / Accepted: 12 February 2025 / Published: 4 March 2025
(This article belongs to the Special Issue Advances in Computational Methods for Fluid Flow)

Abstract

:
Slurry transportation is an essential process in numerous industrial applications, widely studied for its efficiency in material conveyance. Despite substantial research, the impact of pipe wall roughness on critical metrics such as pressure drop, specific energy consumption (SEC), and the Nusselt number remains relatively underexplored. This study provides a detailed analysis using a three-dimensional computational model of a slurry pipeline, with a 0.0549 m diameter and 3.8 m length. The model employs an Eulerian multiphase approach coupled with the RNG k-ε turbulence model, assessing slurry concentrations Cw = 40–60% (by weight). Simulations were conducted at flow velocities Vm = 1–5 m/s, with pipe roughness (Rh) ranging between 10 and 50 µm. Computational findings indicate that both pressure drop and SEC increase proportionally with roughness height, Vm, and Cw. Interestingly, the Nusselt number appears unaffected by roughness height, although it rises corresponds to Vm, and Cw. These insights offer a deeper understanding of slurry pipeline dynamics, informing strategies to enhance operational efficiency and performance across various industrial contexts.

1. Introduction

Slurry transportation through pipelines is a widely utilized method in the chemical, petrochemical, and mining industries, offering efficient transfer of solid particles suspended in a fluid medium, typically water [1]. This method is advantageous due to minimal contamination risk, low maintenance costs, high energy efficiency, and reduced material wastage, making it both cost-effective and sustainable for industrial applications [2]. Extensive research in slurry transportation has deepened our understanding of the complex dynamics and key characteristics involved in pipeline slurry transport.
For example, Kaushal et al. [3] experimentally examined key parameters, including pressure drop and solid concentration profiles for 125 µm, and 440 µm glass beads particles in a straight pipe at Vm = 1–5 m/s, and Cvf = 10–50%. Experimental work shows an increase in pressure drop corresponding with both rise in Vm, and Cvf, with 440 µm particles exhibiting a greater pressure drop than 125 µm particles. Similarly, Joshi et al. [4] investigated the effect of pipe roughness on pressure distribution, with results that pipe roughness marginally impacts pressure drop at low Vm. The optimal transport efficiency achieved by 125 µm particles flowing at Vm = 2 m/s and Cw = 40%. Further studies have expanded on the kinetic and thermodynamic parameters of slurry flow. Gopaliya et al. [5] analyzed various factors such as granular pressure, viscosity, and solid concentration within a straight pipe, highlighting that higher Vm (3.1 m/s), and Cvf (45%) correlate with increased transport dynamics. Wu et al. [6] introduced a model for predicting specific energy consumption (SEC) in low-viscosity Newtonian slurries. He also proposed a method to improve energy efficiency in Newtonian slurry transportation, revealing that energy consumption per unit mass increases with Vm. They also observed that SEC decreases with larger equipment dimensions, such as pipe diameter and tank size. Joshi et al. [7] employed Eulerian–Eulerian approaches in conjunction with the RNG k-ε turbulence model to predict the transportation and efficiency-related parameters of four industrial-oriented solid particles in a straight pipeline: silica sand, iron ore, glass beads, and bottom ash. The sizes for all particles range from 125 to 440 µm, with a velocity of Vm = 1–5 m/s and Cw = 40–60%. The computational investigation shows that smaller solid particles (125–275 µm) flowing at low-velocity Vm = 1–2 m/s and low efflux concentration Cw = 40% have a lower pressure drop and SEC for transportation, indicating that these constraints are highly efficient. Lately, Joshi et al. [8] conducted new research focused on the influence of carrier fluid density on SEC. In their study, an RNG k-ε turbulence multiphase model is used to predict various kinetic parameters, including pressure drop, granular pressure, granular temperature, wall shear stress, settling velocity, and SEC. All parameters are investigated for glass bead particles of five distinct sizes: 125–440 µm, considering different fluid densities and temperatures. Their findings indicated that a carrier fluid with lower density, and higher temperature is more energy-efficient for transportation. Moreover, they observed that the characteristics of slurry flow were influenced by particle size and solid concentration. Specifically, larger solid particles at higher solid concentrations were found to have higher settling velocities but lower transport efficiency.
Joshi et al. [9] used the Eulerian k-ε multiphase model in one of the early studies centered around a bi-modal flow in a straight pipe. The study is carried out for four distinct silica sand and fly ash mixture combinations, flowing at Vm = 2–5 m/s, and Cw = 40–60%. According to the findings, the 65:35 combination has the lowest pressure drop, which rises with Vm and Cvf. The results also show that slurry with a high percentage of catalyst particles, i.e., a 65:35 mix, improves slurry system transport efficiency. Moving forward to the heat transfer characteristics, Nayak et al. [10] investigated the heat transfer characteristics of the fly ash–water slurry mixture using the k-ε turbulence model, flowing through a horizontal pipeline with an inlet temperature of 300 K and wall temperature of 400 °K. The investigation uses five different particle sizes (4–78 µm), with a Vm = 1–5 m/s, and a Cvf = 10–50%. They observed that while the heat transfer coefficient increased with Vm, it decreased with higher Cvf. Additionally, larger particles yielded the highest heat transfer ratio. Moreover, the Nusselt number increases with a rise in Reynolds number.
The literature reveals that only a few studies are available on pipe roughness and their influence on pressure drop and concentration distribution. However, their effect on transport efficiency parameter (SEC) and heat transfer parameter (Nusselt number) has not yet been established. Therefore, the present research utilizes the RNG k-ε turbulence model and granular flow theory to evaluate the effects of pipe roughness on transport (pressure gradient and SEC) and thermal (Nusselt number) characteristics in a mono-dispersed glass bead slurry. Flow conditions include Vm = 1–5 m/s, and Cw = 40–60%. Computational results were validated against experimental data, providing insights for optimizing pipeline design in industries like petroleum, mineral processing, and chemicals reliant on slurry transportation.

2. Mathematical Formulation

2.1. Governing Equations

The numerical analysis employs the Eulerian multiphase model, which separately solves conservation equations for mass, momentum, and energy for each distinct phase in the system. This approach allows for an accurate representation of interactions between the solid and fluid phases. For comprehensive formulations of the continuity and momentum equations applied to both phases, a detailed description of the equations can be found in Joshi et al. [7] and Nayak et al. [10].

2.2. Heat Transfer Coefficient

The heat transfer rate, which takes place between multiple phases, can be calculated as a function of the temperature difference. The equation can be written as [10]:
Q s f = h s f T s T f
where h s f = h f s is the heat transfer coefficient that takes place between the solid and fluidic phases.
The correlation of the Nusselt number used by Gun et al. [11] for granular flow can be written as:
N u s = 7 10 α f + 5 α f 2 1 + 0.7 Re s 0.2 Pr 0.33 + 1.33 2.4 α f + 1.2 α f 2 Re s 0.7 Pr 0.33
Here, α f is the volume fraction of the fluid phase, R e s   and Pr are represented as the relative Reynolds number and fluid Prandtl number.

2.3. Specific Energy Consumption (SEC) Equation

SEC for the slurry transportation model is termed as the minimum amount of required energy for the solid particle transportation to the unit distance throughout the pipeline from inlet to outlet. The SEC is the function of pressure gradient, solid density, and volumetric concentration, which can be expressed as follows [6]:
S E C = Δ P Δ L C V ρ S

3. Computational Modeling

Numerical modeling is a robust method to accurately forecast crucial multiphase factors. The primary focus of modeling orbits around solving the partial differential equations pertaining to momentum and energy. The commercial CFD software Ansys Fluent 2024R1 (ANSYS Inc., Canonsburg, PA, USA) is employed to analyze the kinetic and thermal characteristics of a slurry mixture flowing through a horizontal, straight pipeline [12,13,14]. This multiphase mono-dispersed model encompasses mainly two phases, (i) solid (glass beads), and (ii) liquid (water). Previous researchers have effectively used the Eulerian multiphase k-ε model to predict the kinetic and thermal behaviors of slurry flow. In present work, a straight pipe, oriented horizontally, measuring L = 3.8 m and D = 0.0549 m, serves as the conduit for transporting the mixture from inlet to the outlet. The chosen pipe geometry satisfies the criteria for fully developed flow, i.e., 50D. The schematic diagram of the computational model, which includes pipe geometry and boundary conditions, is depicted in Figure 1.

3.1. Meshed Model

Figure 2a displays the 3D meshed model of a straight pipe. In this meshing process, hexahedral elements are employed to create the mesh for the pipe model. The computational model exhibits an excellent orthogonal quality of 0.93417, which is highly desirable for CFD simulations, ensuring precise results [15]. Additionally, the skewness of the pipe is maintained at a low value of 0.0256, which adheres to recommended standards for CFD models. To accurately model the interaction between particles and the pipe wall, the “near-wall” mesh was suitably refined, ensuring a y+ value of 1 to account for boundary layer effects.
To maintain a consistent mesh quality and enhance computational efficiency, a mesh independence test was carried out. Multiple mesh configurations were generated, including element counts of 210 × 103–661 × 103, as depicted in Figure 2b,c. After careful examination, it was noted that the computational models with 472 × 103 and 661 × 103 elements produced virtually indistinguishable velocity distribution results, as shown in Figure 2b. In a similar way, analogous mesh elements are accounted for to obtain the pressure drop. The results show similar patterns, i.e., with 472 × 103 and 661 × 103 element mesh shows approximately similar results, see Figure 2c. As a result, the pipe geometry utilizing 472 × 103 elements was chosen as the optimal setup for conducting the simulations.

3.2. Boundary Condition and Convergence Criteria

A slurry mixture consists of solid particles and water flows at a specific velocity (Vm = 1–5 m/s) and volumetric concentrations (Cw = 40–60%) from the pipe inlet section. The mean diameter and specific gravity of solid particles are considered as 125 µm and 2.470 Kg/m3, respectively. The no-slip condition is used for the pipeline wall [9] to obtain the kinetic and heat transfer characteristics [16]. An Eulerian technique combined with the RNG k-ε turbulent model is employed to simulate mono-dispersed slurry flow for a continuous fluid phase [17]. To achieve accurate and smooth characteristics in the slurry flow, a combination of a second-order upwind scheme and a SIMPLE algorithm with a convergence criterion set at 10−4 is employed.

3.3. Numerical Validation

A comparison of numerically obtained results and experimental results for 125 μm solid particles is illustrated in Figure 3. The comparison is carried out for glass bead particles flowing at Vm = 1–5 m/s and efflux concentration Cvf = 30%, and Cvf = 50%, see Figure 3a,b, respectively. From Figure 3, it can be seen that the computational outcome for solid particles is constant with the experimental work by Kaushal et al. [3].

4. Results and Discussion

4.1. Pressure Distribution

The pressure exerted by the slurry mixture results from the interactions between particles and the carrier fluid, causing a frictional pressure drop. Figure 4a–c depicts the normalized pressure drop for 125 µm solid particles as a function of pipe roughness (roughness height) while varying the Cw from 40% to 60%. The x-axis represents the normalized pressure drop, and the y-axis spans roughness heights of 10 to 50 µm.
As seen in Figure 4a, the findings demonstrate a direct relationship between pressure drop and roughness height. Slurry flowing through pipes with lower roughness heights exhibits less pressure drop, while the pressure drop is highest when the roughness height is 50 µm. Furthermore, Vm has a substantial impact on pressure drop, with an increase in slurry Vm leading to higher pressure drop. At lower Vm (1–2 m/s), the rise in pressure drop experienced due to increased pipe roughness is relatively moderate. However, at Vm exceeding the 2 m/s mark, a significant surge in pressure drop is experienced, especially noticeable in the Vm = 3–5 m/s. This surge in pressure drop is attributed to the elevated friction factors encountered at higher slurry flow Vm within the pipeline.
Cw also plays a noteworthy role in pressure drop, as shown in Figure 4b,c. The literature confirms that an increase in Cw results in a corresponding rise in pressure drop for various roughness heights and Vm. This can be explained by the fact that higher concentrations prevent particles from settling out prematurely, leading to increased interactions between solid particles and between solid and liquid components, consequently causing an increase in pressure drop.

4.2. Concentration Distribution

The figures depict the concentration distribution of 125 µm particles transported through a pipe with varying surface roughness heights (Rh = 10 µm, 30 µm, and 50 µm) at different Vm = 1 m/s (Figure 5a), 3 m/s (Figure 5b), and 5 m/s (Figure 5c). The color contour maps indicate particle concentration levels across the pipe cross-section. At low Vm (1 m/s), the particle concentration near the wall shows a more uneven distribution, with a visible impact of surface roughness on particle accumulation patterns, particularly for higher roughness heights. As Vm increases to 3 m/s, the distribution becomes more homogenized in the central region, although the influence of surface roughness persists near the wall. At the highest Vm = 5 m/s, particle concentration in the core flow region becomes nearly uniform, suggesting dominant inertial effects, while the wall regions still exhibit roughness-induced gradients. The roughness impact is mainly shown at lower Vm, whereas higher Vm reduces the relative impact of roughness due to stronger bulk flow and particle inertia.
Figure 6 illustrates the concentration distribution of 125 µm particles transported through a pipe at a velocity of Vm = 5 m/s, and Rh = 50 µm, with varying Cw = 40–60%. As the Cw increases, a pronounced rise in overall particle concentration is observed across the cross-section. At Cw = 40%, the particle distribution is relatively uniform, with slightly higher concentrations near the pipe’s lower regions. For Cw = 50%, the concentration becomes more prominent and evenly distributed across the core flow area, reflecting a moderate increase in particle density. At Cw = 60%, the particle concentration reaches its highest levels, as indicated by the dominance of orange and red shades, especially in the central and near-wall regions.

4.3. Specific Energy Consumption (SEC)

The efficiency of the slurry pipeline system’s material transportation is quantified by a metric known as SEC. Figure 7a–c displays the normalized SEC trends for 125 µm solid particles concerning pipe roughness (Rh) at Cw = 40–60%. The x-axis represents the normalized SEC, while the y-axis spans Rh from 10 to 50 µm.
For Cw = 40%, it is evident that energy consumption increases as Rh rises. This is attributed to the elevated Rh creating an additional frictional barrier, hindering slurry flow within the pipeline system. Consequently, pressure drop increases. The impact of Vm on SEC is noteworthy. The SEC of the slurry system escalates with a higher slurry Vm. At lower Vm (1–2 m/s), minimal deviation is observed in the pressure drop trendline. However, as Vm surpasses 2 m/s, a substantial rise in SEC becomes apparent, particularly in the Vm range of 3 to 5 m/s. This increase in SEC is linked to the elevated pressure drop encountered at higher Vm within the slurry pipeline.
Cw also significantly influences SEC, as depicted in Figure 7b,c. The literature confirms that an increase in Cw leads to a corresponding rise in SEC for various Rh and Vm. This is due to higher concentrations resulting in greater pressure drop, thereby requiring more power for transportation, consequently leading to an increase in SEC.

4.4. Nusselt Number

The Nusselt number, a dimensionless parameter commonly used to assess heat transfer between a moving fluid and a solid surface, is examined in Figure 8. This illustration displays the relationship between the Nusselt number and the average Vm, specifically for particles sized at 125 µm. The graph provides Nusselt number values for three distinct Cw = 40–60%, transporting at Vm = 1 to 5 m/s. The chart highlights the positive correlation between the Nusselt number and the mixture Vm. Slurry flowing at lower Vm exhibits lower Nusselt numbers, which progressively increase with the mixture Vm. This behavior is attributed to higher flow Vm enhancing heat transfer between the solid and liquid phases. Furthermore, it is noteworthy that slurry mixtures with lower Cw = 40% display the lowest Nusselt numbers, while those with Cw = 60% exhibit the highest values. This variation is due to the increased heat transferred by the higher-concentration slurry mixture.

5. Conclusions

This study investigates the influence of pipe roughness height (Rh) on the kinetic and thermo-fluidic transport characteristics of a solid–liquid slurry flowing through a horizontal straight pipe. An Eulerian multiphase model combined with the RNG k-ε turbulence model was employed to analyze the slurry’s kinetic and thermal behavior. Key findings from the numerical analysis include:
  • The Eulerian model accurately predicts pressure drop for 125 µm particles in a horizontal straight pipe, with Vm = 1–5 m/s, and Cvf = 30%, showing strong agreement with experimental data.
  • The normalized pressure gradient positively correlates with roughness height, which further varies with Vm: minimal at low Vm but increasing exponentially at higher Vm. Additionally, for a given Vm and roughness height, the pressure gradient increases proportionally with Cw.
  • The surface roughness significantly affects particle concentration distribution at lower velocities, with roughness-induced gradients diminishing as Vm increases due to stronger inertial effects. Additionally, higher Cw results in a more uniform yet denser particle distribution across the pipe cross-section.
  • Similarly to the pressure gradient, normalized Specific Energy Consumption (SEC) also rises with roughness height, Vm, and Cw, following a comparable trend.
  • The non-dimensional Nusselt number remains independent of roughness height yet is significantly influenced by Vm and Cw. A lower Vm corresponds to reduced Nusselt numbers, which increase with arh rise in Vm and Cw. The maximum Nusselt number is observed at Cw = 60%.
Based on these findings, reducing pipe roughness height is recommended to enhance system efficiency and decrease additional power requirements.

Author Contributions

Conceptualization, T.J.; methodology, T.J., O.P. and G.K.; software, T.J.; validation, T.J. and G.K.; formal analysis, T.J. and O.P.; investigation, T.J. and G.K.; resources, G.K. and R.K.B.G.; data curation, G.K. and R.K.B.G.; writing—original draft preparation, T.J. and G.K.; writing—review and editing, R.K.B.G. and G.K.; visualization, T.J.; supervision, G.K. and O.P.; project administration, G.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data will be available on request.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

C v f Solid concentration (% by volume) C w Solid concentration (% by weight)
D Diameter of Pipe (m)hHeat transfer coefficient (W/m2K)
L Length of pipe (m) N u s Nusselt number
PrPrandtl number Δ P / Δ L Pressure Gradient (Pa/m)
Q s f Interphase heat transfer term (W/m3) R e s Relative Reynolds number
RhPipe roughness or Roughness height T f Fluid Temperature (°K)
T s Solid Temperature (K) V m Mean flow velocity (m/s)
α f Volume fraction by fluid phase ρ s Density of solid phase (kg/m3)

References

  1. Joshi, T.; Parkash, O.; Krishan, G. CFD modeling for slurry flow through a horizontal pipe bend at different Prandtl number. Int. J. Hydrogen Energy 2022, 47, 23731–23750. [Google Scholar] [CrossRef]
  2. Joshi, T.; Parkash, O.; Gallegos, R.K.B.; Krishan, G. Computational investigation of transportation and thermal characteristics in a bi-modal slurry flow through a horizontally placed pipe bend. Powder Technol. 2024, 442, 119879. [Google Scholar] [CrossRef]
  3. Kaushal, D.R.; Tomita, Y. Experimental investigation for near-wall lift of coarser particles in slurry pipeline using γ-ray densitometer. Powder Technol. 2007, 172, 177–187. [Google Scholar] [CrossRef]
  4. Joshi, T.; Parkash, O.; Krishan, G. Numerical Investigation of Slurry Pressure Drop at Different Pipe Roughness in a Straight Pipe Using CFD. Arab. J. Sci. Eng. 2022, 47, 15391–15414. [Google Scholar] [CrossRef]
  5. Gopaliya, M.K.; Kaushal, D.R. Analysis of Effect of Grain Size on Various Parameters of Slurry Flow through Pipeline Using CFD. Part. Sci. Technol. 2015, 33, 369–384. [Google Scholar] [CrossRef]
  6. Wu, J.; Graham, L.; Wang, S.; Parthasarathy, R. Energy efficient slurry holding and transport. Miner. Eng. 2010, 23, 705–712. [Google Scholar] [CrossRef]
  7. Joshi, T.; Parkash, O.; Krishan, G. Estimation of energy consumption and transportation characteristics for slurry flow through a horizontal straight pipe using computational fluid dynamics. Phys. Fluids 2023, 35, 053303. [Google Scholar] [CrossRef]
  8. Joshi, T.; Parkash, O.; Krishan, G. Slurry flow characteristics through a horizontal pipeline at different Prandtl number. Powder Technol. 2022, 413, 118008. [Google Scholar] [CrossRef]
  9. Joshi, T.; Parkash, O.; Murthy, A.A.; Krishan, G. Numerical investigation of Bi-model slurry transportation in a straight pipe. Results Eng. 2023, 17, 100858. [Google Scholar] [CrossRef]
  10. Nayak, B.B.; Chatterjee, D. Numerical investigation of convective heat transfer in pipeline flow of multi-sized mono dispersed fly ash-water slurry. Int. J. Heat Mass Transf. 2017, 108, 1802–1818. [Google Scholar] [CrossRef]
  11. Gunn, D.J. Transfer of heat or mass to particles in fixed and fluidised beds. Int. J. Heat Mass Transf. 1978, 21, 467–476. [Google Scholar] [CrossRef]
  12. Razavi, F.; Komrakova, A.; Lange, C.F. CFD–DEM Simulation of Sand-Retention Mechanisms in Slurry Flow. Energies 2021, 14, 3797. [Google Scholar] [CrossRef]
  13. Shi, H.; Yuan, J.; Li, Y. The Impact of Swirls on Slurry Flows in Horizontal Pipelines. J. Mar. Sci. Eng. 2021, 9, 1201. [Google Scholar] [CrossRef]
  14. Joshi, T.; Parkash, O.; Gallegos, R.K.B.; Krishan, G. Comparative investigation of the energy consumption and heat transfer characteristics of Uni-modal and Bi-modal slurry flow through a straight pipe. Ocean Eng. 2025, 316, 119987. [Google Scholar] [CrossRef]
  15. Joshi, T.; Gupta, A.; Parkash, O.; Gallegos, R.K.B.; Oo, N.L.; Krishan, G. Mathematical Formulations for Predicting Pressure Drop in Solid–Liquid Slurry Flow through a Straight Pipe Using Computational Modeling. Mathematics 2024, 12, 2879. [Google Scholar] [CrossRef]
  16. Li, H.; Prakash, A. Heat Transfer and Hydrodynamics in a Three-Phase Slurry Bubble Column. Ind. Eng. Chem. Res. 1997, 36, 4688–4694. [Google Scholar] [CrossRef]
  17. Kaushal, D.R.; Kumar, A.; Tomita, Y.; Kuchii, S.; Tsukamoto, H. Flow of mono-dispersed particles through horizontal bend. Int. J. Multiph. Flow 2013, 52, 71–91. [Google Scholar] [CrossRef]
Figure 1. Schematic Diagram of Computational Flow Domain.
Figure 1. Schematic Diagram of Computational Flow Domain.
Computation 13 00065 g001
Figure 2. (a) Computational meshed model, Mesh independency test for (b) Velocity distribution, and (c) Pressure gradient distribution.
Figure 2. (a) Computational meshed model, Mesh independency test for (b) Velocity distribution, and (c) Pressure gradient distribution.
Computation 13 00065 g002aComputation 13 00065 g002b
Figure 3. Numerical validation of 125 µm solid particle with Kaushal et al. [3] for, (a) Cvf = 30%, and (b) Cvf = 50%.
Figure 3. Numerical validation of 125 µm solid particle with Kaushal et al. [3] for, (a) Cvf = 30%, and (b) Cvf = 50%.
Computation 13 00065 g003
Figure 4. Normalized pressure gradient for 125 µm particle transported through pipe with different roughness height at efflux concentration, (a) Cw = 40%, (b) Cw = 50%, and (c) Cw = 60%.
Figure 4. Normalized pressure gradient for 125 µm particle transported through pipe with different roughness height at efflux concentration, (a) Cw = 40%, (b) Cw = 50%, and (c) Cw = 60%.
Computation 13 00065 g004aComputation 13 00065 g004b
Figure 5. Concentration distribution for 125 µm particle transported through pipe with different roughness height at velocity, (a) Vm = 1 m/s, (b) Vm = 3 m/s, and (c) Vm = 5 m/s.
Figure 5. Concentration distribution for 125 µm particle transported through pipe with different roughness height at velocity, (a) Vm = 1 m/s, (b) Vm = 3 m/s, and (c) Vm = 5 m/s.
Computation 13 00065 g005
Figure 6. Concentration distribution for 125 µm particle transported through pipe with Vm = 5 m/s, Rh = 50 µm at efflux concentration Cw = 40–60%.
Figure 6. Concentration distribution for 125 µm particle transported through pipe with Vm = 5 m/s, Rh = 50 µm at efflux concentration Cw = 40–60%.
Computation 13 00065 g006
Figure 7. Normalized specific energy consumption for a 125 µm particle transported through pipe with different roughness height at efflux concentration, (a) Cw = 40%, (b) Cw = 50%, and (c) Cw = 60%.
Figure 7. Normalized specific energy consumption for a 125 µm particle transported through pipe with different roughness height at efflux concentration, (a) Cw = 40%, (b) Cw = 50%, and (c) Cw = 60%.
Computation 13 00065 g007aComputation 13 00065 g007b
Figure 8. Variation in Nusselt number for Cw = 40–60% flowing at mixture velocity Vm = 1–5 m/s.
Figure 8. Variation in Nusselt number for Cw = 40–60% flowing at mixture velocity Vm = 1–5 m/s.
Computation 13 00065 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Joshi, T.; Parkash, O.; Gallegos, R.K.B.; Krishan, G. Computational Analysis of Pipe Roughness Influence on Slurry Flow Dynamics. Computation 2025, 13, 65. https://doi.org/10.3390/computation13030065

AMA Style

Joshi T, Parkash O, Gallegos RKB, Krishan G. Computational Analysis of Pipe Roughness Influence on Slurry Flow Dynamics. Computation. 2025; 13(3):65. https://doi.org/10.3390/computation13030065

Chicago/Turabian Style

Joshi, Tanuj, Om Parkash, Ralph Kristoffer B. Gallegos, and Gopal Krishan. 2025. "Computational Analysis of Pipe Roughness Influence on Slurry Flow Dynamics" Computation 13, no. 3: 65. https://doi.org/10.3390/computation13030065

APA Style

Joshi, T., Parkash, O., Gallegos, R. K. B., & Krishan, G. (2025). Computational Analysis of Pipe Roughness Influence on Slurry Flow Dynamics. Computation, 13(3), 65. https://doi.org/10.3390/computation13030065

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