[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (402)

Search Parameters:
Keywords = Newtonian flow model

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 12159 KiB  
Article
Numerical Study of Carreau Fluid Flow in Symmetrically Branched Tubes
by Vinicius Pepe, Antonio F. Miguel, Flávia Zinani and Luiz Rocha
Symmetry 2025, 17(1), 48; https://doi.org/10.3390/sym17010048 - 30 Dec 2024
Viewed by 256
Abstract
The non-Newtonian Carreau fluid model is a suitable model for pseudoplastic fluids and can be used to characterize fluids not so different from biological fluids, such as the blood, and fluids involved in geological processes, such as lava and magma. These fluids are [...] Read more.
The non-Newtonian Carreau fluid model is a suitable model for pseudoplastic fluids and can be used to characterize fluids not so different from biological fluids, such as the blood, and fluids involved in geological processes, such as lava and magma. These fluids are frequently conveyed by complex flow structures, which consist of a network of channels that allow the fluid to flow from one place (source or sink) to a variety of locations or vice versa. These flow networks are not randomly arranged but show self-similarity at different spatial scales. Our work focuses on the design of self-similar branched flow networks that look the same on any scale. The flow is incompressible and stationary with a viscosity following the Carreau model, which is important for the study of complex flow systems. The flow division ratios, the flow resistances at different scales, and the geometric size ratios for maximum flow access are studied, based on Computational Fluid Dynamics (CFD). A special emphasis is placed on investigating the possible incidence of flow asymmetry in these symmetric networks. Our results show that asymmetries may occur for both Newtonian and non-Newtonian fluids and shear-thinning fluids most affect performance results. The lowest flow resistance occurs when the diameters of the parent and daughter ducts are equal, and the more uniform distribution of flow resistance occurs for a ratio between the diameters of the parent and daughter ducts equal to 0.75. Resistances for non-Newtonian fluids are 4.8 to 5.6 times greater than for Newtonian fluids at Reynolds numbers of 100 and 250, respectively. For the design of engineering systems and the assessment of biological systems, it is recommended that the findings presented are taken into account. Full article
(This article belongs to the Special Issue Symmetry in Thermal Fluid Sciences and Energy Applications)
Show Figures

Figure 1

Figure 1
<p>Self-similar fluidic structure with three symmetric branching levels and circular sections to transport Carreau’s fluids.</p>
Full article ">Figure 2
<p>Contours of static pressure for a network designed for a<sub>D</sub> = 0.80.</p>
Full article ">Figure 2 Cont.
<p>Contours of static pressure for a network designed for a<sub>D</sub> = 0.80.</p>
Full article ">Figure 3
<p>Dimensionless flow resistance of a Carreau fluid (Equation (18)) versus the diameter ratio a<sub>D</sub> and svelteness index Sv, for Reynolds numbers 100 and 250.</p>
Full article ">Figure 3 Cont.
<p>Dimensionless flow resistance of a Carreau fluid (Equation (18)) versus the diameter ratio a<sub>D</sub> and svelteness index Sv, for Reynolds numbers 100 and 250.</p>
Full article ">Figure 4
<p>CNR flow resistances (Equation (19)) versus the diameter ratio a<sub>D</sub> and svelteness index Sv, for Reynolds numbers 100 and 250.</p>
Full article ">Figure 5
<p>Euler number (Equation (17)) versus the diameter ratio a<sub>D</sub>, svelteness index Sv, Reynolds number, Carreau number, and rheological parameters.</p>
Full article ">Figure 6
<p>Dimensionless flow resistance (R<sub>i</sub>/R<sub>T</sub>) versus diameter ratio for different Reynolds numbers and rheological parameters.</p>
Full article ">Figure 7
<p>Dimensionless flow resistance (R<sub>i</sub>/R<sub>T</sub>) for best resistance distributions, versus the diameter ratio a<sub>D</sub> and svelteness index Sv.</p>
Full article ">Figure 8
<p>Dimensionless flow partitioning ratio (Equation (20)) for level 1.</p>
Full article ">Figure 9
<p>Dimensionless flow partitioning ratio (Equation (20)) for level 2.</p>
Full article ">Figure 10
<p>Dimensionless flow partitioning ratio (Equation (20)) for level 3.</p>
Full article ">
18 pages, 4698 KiB  
Article
Computational Fluid Dynamics Simulation and Analysis of Non-Newtonian Drilling Fluid Flow and Cuttings Transport in an Eccentric Annulus
by Muhammad Ahsan, Shah Fahad and Muhammad Shoaib Butt
Mathematics 2025, 13(1), 101; https://doi.org/10.3390/math13010101 - 30 Dec 2024
Viewed by 282
Abstract
This study examines the flow behavior as well as the cuttings transport of non-Newtonian drilling fluid in the geometry of an eccentric annulus, accounting for what impacts drill pipe rotation might have on fluid velocity, as well as annular eccentricity on axial and [...] Read more.
This study examines the flow behavior as well as the cuttings transport of non-Newtonian drilling fluid in the geometry of an eccentric annulus, accounting for what impacts drill pipe rotation might have on fluid velocity, as well as annular eccentricity on axial and tangential distributions of velocity. A two-phase Eulerian–Eulerian model was developed by using computational fluid dynamics to simulate drilling fluid flow and cuttings transport. The kinetic theory of granular flow was used to study the dynamics and interactions of cuttings transport. Non-Newtonian fluid properties were modeled using power law and Bingham plastic formulations. The simulation results demonstrated a marked improvement in efficiency, as much as 45%, in transport by increasing the fluid inlet velocity from 0.54 m/s to 2.76 m/s, reducing the amount of particle accumulation and changing axial and tangential velocity profiles dramatically, particularly at narrow annular gaps. At a 300 rpm rotation, the drill pipe brought on a spiral flow pattern, which penetrated tangential velocities in the narrow gap that had increased transport efficiency to almost 30% more. Shear-thinning behavior characterizes fluid of which the viscosity, at nearly 50% that of the central core low-shear regions, was closer to the wall high-shear regions. Fluid velocity and drill pipe rotation play a crucial role in optimizing cuttings transport. Higher fluid velocities with controlled drill pipe rotation enhance cuttings removal and prevent particle build-up, thereby giving very useful guidance on how to clean the wellbore efficiently in drilling operations. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the eccentric annulus.</p>
Full article ">Figure 2
<p>Computational mesh and geometry domain.</p>
Full article ">Figure 3
<p>Axial velocities: (<b>a</b>) Re 1140, 0 rmp, Plane-1; (<b>b</b>) 1150, 300 rmp, Plane-1; (<b>c</b>) Re 1140, 0 rmp, Plane-2; (<b>d</b>) 1150, 300 rmp, Plane-2; (<b>e</b>) Re 1140, 0 rmp, Plane-3; (<b>f</b>) 1150, 300 rmp, Plane-3 (literature data source [<a href="#B23-mathematics-13-00101" class="html-bibr">23</a>]).</p>
Full article ">Figure 4
<p>Axial velocities: (<b>a</b>) Re 9300, 0 rmp, Plane-1; (<b>b</b>) 9200, 300 rmp, Plane-1; (<b>c</b>) Re 9300, 0 RPMrmpPlane-2; (<b>d</b>) 9200, 300 rmp, Plane-2; (<b>e</b>) Re 9300, 0 RPM, Plane-3; (<b>f</b>) 9200, 300 rmp, Plane-3 (literature data source [<a href="#B23-mathematics-13-00101" class="html-bibr">23</a>]).</p>
Full article ">Figure 5
<p>Tangential velocities: (<b>a</b>) Re 1150, 300 rpm, Plane-1; (<b>b</b>) 9200, 300 rpm, Plane-1; (<b>c</b>) Re 1150, 300 rpm, Plane-2; (<b>d</b>) 9200, 300 rpm, Plane-2; (<b>e</b>) Re 1150, 300 rpm, Plane-3; (<b>f</b>) 9200, 300 rpm, Plane-3 (literature data source [<a href="#B23-mathematics-13-00101" class="html-bibr">23</a>]).</p>
Full article ">Figure 6
<p>Velocity magnitude contours: (<b>a</b>) Re: 1140, 0 rpm; (<b>b</b>) Re: 1150, 300 rpm; (<b>c</b>) Re: 9300, 0 rpm; (<b>d</b>) Re: 9200, 300 rpm.</p>
Full article ">Figure 7
<p>Velocity vectors: (<b>a</b>) Re: 1140, 0 rpm; (<b>b</b>) Re: 1150, 300 rpm; (<b>c</b>) Re: 9300, 0 rpm; (<b>d</b>) Re: 9200, 300 rpm.</p>
Full article ">Figure 8
<p>Molecular viscosity contours: (<b>a</b>) Re: 1140, 0 rpm; (<b>b</b>) Re: 1150, 300 rpm; (<b>c</b>) Re: 9300, 0 rpm; (<b>d</b>) Re: 9200, 300 rpm.</p>
Full article ">
12 pages, 2722 KiB  
Article
Impact of Addition of a Newtonian Solvent to a Giesekus Fluid: Analytical Determination of Flow Rate in Plane Laminar Motion
by Irene Daprà, Giambattista Scarpi and Vittorio Di Federico
Fluids 2025, 10(1), 1; https://doi.org/10.3390/fluids10010001 - 24 Dec 2024
Viewed by 270
Abstract
In this study, the influence of the presence of a Newtonian solvent on the flow of a Giesekus fluid in a plane channel or fracture is investigated with a focus on the determination of the flow rate for an assigned external pressure gradient. [...] Read more.
In this study, the influence of the presence of a Newtonian solvent on the flow of a Giesekus fluid in a plane channel or fracture is investigated with a focus on the determination of the flow rate for an assigned external pressure gradient. The pressure field is nonlinear due to the presence of the normal transverse stress component. As expected, the flow rate per unit width Q is larger than for a Newtonian fluid and decreases as the solvent increases. It is strongly dependent on the viscosity ratio ε (0ε1), the dimensionless mobility parameter β (0β1) and the Deborah number De, the dimensionless driving pressure gradient. The degree of dependency is notably strong in the low range of ε. Furthermore, Q increases with De and tends to a constant asymptotic value for large De, subject to the limitation of laminar flow. When the mobility factor β is in the range 0.5÷1, there is a minimum value of ε  to obtain an assigned value of De. The ratio UN/U between Newtonian and actual mean velocity depends only on the product βDe, as for other non-Newtonian fluids. Full article
(This article belongs to the Special Issue Advances in Computational Mechanics of Non-Newtonian Fluids)
Show Figures

Figure 1

Figure 1
<p>Flow configuration.</p>
Full article ">Figure 2
<p>The flow rate versus Deborah number: (<b>a</b>) for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>b</b>) for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.75</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Variation of the flow rate versus Deborah number for some <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>&lt;</mo> <mn>0.5</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Ratio between the two mean velocities versus elasticity with <math display="inline"><semantics> <mrow> <mi>ε</mi> </mrow> </semantics></math> as a parameter: (<b>a</b>) for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>; (<b>b</b>) for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.75</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Ratio between the maximum velocity and the mean velocity versus <math display="inline"><semantics> <mrow> <mi>ε</mi> </mrow> </semantics></math> for some values of <math display="inline"><semantics> <mrow> <mi>D</mi> <mi>e</mi> <mo>,</mo> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>.</p>
Full article ">
12 pages, 575 KiB  
Article
Symmetry Analysis of the 3D Boundary-Layer Flow of a Non-Newtonian Fluid
by Ali El Saheli and Bashar Zogheib
AppliedMath 2024, 4(4), 1588-1599; https://doi.org/10.3390/appliedmath4040084 - 20 Dec 2024
Viewed by 358
Abstract
This study investigates the three-dimensional, steady, laminar boundary-layer equations of a non-Newtonian fluid over a flat plate in the absence of body forces. The classical boundary-layer theory, introduced by Prandtl in 1904, suggests that fluid flows past a solid surface can be divided [...] Read more.
This study investigates the three-dimensional, steady, laminar boundary-layer equations of a non-Newtonian fluid over a flat plate in the absence of body forces. The classical boundary-layer theory, introduced by Prandtl in 1904, suggests that fluid flows past a solid surface can be divided into two regions: a thin boundary layer near the surface, where steep velocity gradients and significant frictional effects dominate, and the outer region, where friction is negligible. Within the boundary layer, the velocity increases sharply from zero at the surface to the freestream value at the outer edge. The boundary-layer approximation significantly simplifies the Navier–Stokes equations within the boundary layer, while outside this layer, the flow is considered inviscid, resulting in even simpler equations. The viscoelastic properties of the fluid are modeled using the Rivlin–Ericksen tensors. Lie group analysis is applied to reduce the resulting third-order nonlinear system of partial differential equations to a system of ordinary differential equations. Finally, we determine the admissible forms of the freestream velocities in the x- and z-directions. Full article
Show Figures

Figure 1

Figure 1
<p>Laminar boundary layer.</p>
Full article ">
36 pages, 1641 KiB  
Review
The Reynolds Number: A Journey from Its Origin to Modern Applications
by Manuel Saldana, Sandra Gallegos, Edelmira Gálvez, Jonathan Castillo, Eleazar Salinas-Rodríguez, Eduardo Cerecedo-Sáenz, Juan Hernández-Ávila, Alessandro Navarra and Norman Toro
Fluids 2024, 9(12), 299; https://doi.org/10.3390/fluids9120299 - 16 Dec 2024
Viewed by 545
Abstract
The Reynolds number (Re), introduced in the late 19th century, has become a fundamental parameter in a lot of scientific fields—the main one being fluid mechanics—as it allows for the determination of flow characteristics by distinguishing between laminar and turbulent regimes, or some [...] Read more.
The Reynolds number (Re), introduced in the late 19th century, has become a fundamental parameter in a lot of scientific fields—the main one being fluid mechanics—as it allows for the determination of flow characteristics by distinguishing between laminar and turbulent regimes, or some intermediate stage. Reynolds’ 1895 paper, which decomposed velocity into average and fluctuating components, laid the foundation for modern turbulence modeling. Since then, the concept has been applied to various fields, including external flows—the science that studies friction—as well as wear, lubrication, and heat transfer. Literature research in recent times has explored new interpretations of Re, and despite its apparent simplicity, the precise prediction of Reynolds numbers remains a computational challenge, especially under conditions such as the study of multiphase flows, non-Newtonian fluids, highly turbulent flow conditions, flows on very small scales or nanofluids, flows with complex geometries, transient or non-stationary flows, and flows of fluids with variable properties. Reynolds’ work, which encompasses both scientific and engineering contributions, continues to influence research and applications in fluid dynamics. Full article
(This article belongs to the Special Issue Recent Advances in Fluid Mechanics: Feature Papers, 2024)
Show Figures

Figure 1

Figure 1
<p>Distribution of documents by type (<b>a</b>), field of study (<b>b</b>), and temporal variation (<b>c</b>) with respect to the Reynolds number in the SCOPUS bibliographic reference database.</p>
Full article ">Figure 1 Cont.
<p>Distribution of documents by type (<b>a</b>), field of study (<b>b</b>), and temporal variation (<b>c</b>) with respect to the Reynolds number in the SCOPUS bibliographic reference database.</p>
Full article ">Figure 2
<p>Diagrams of laminar (<b>a</b>) and turbulent (<b>b</b>) flow regimes.</p>
Full article ">
22 pages, 506 KiB  
Article
Topological Degree for Operators of Class (S)+ with Set-Valued Perturbations and Its New Applications
by Evgenii S. Baranovskii and Mikhail A. Artemov
Fractal Fract. 2024, 8(12), 738; https://doi.org/10.3390/fractalfract8120738 - 14 Dec 2024
Viewed by 471
Abstract
We investigate the topological degree for generalized monotone operators of class (S)+ with compact set-valued perturbations. It is assumed that perturbations can be represented as the composition of a continuous single-valued mapping and an upper semicontinuous set-valued mapping with aspheric [...] Read more.
We investigate the topological degree for generalized monotone operators of class (S)+ with compact set-valued perturbations. It is assumed that perturbations can be represented as the composition of a continuous single-valued mapping and an upper semicontinuous set-valued mapping with aspheric values. This allows us to extend the standard degree theory for convex-valued operators to set-valued mappings whose values can have complex geometry. Several theoretical aspects concerning the definition and main properties of the topological degree for such set-valued mappings are discussed. In particular, it is shown that the introduced degree has the homotopy invariance property and can be used as a convenient tool in checking the existence of solutions to corresponding operator inclusions. To illustrate the applicability of our approach to studying models of real processes, we consider an optimal feedback control problem for the steady-state internal flow of a generalized Newtonian fluid in a 3D (or 2D) bounded domain with a Lipschitz boundary. By using the proposed topological degree method, we prove the solvability of this problem in the weak formulation. Full article
(This article belongs to the Special Issue Fixed Point Theory and Fractals)
Show Figures

Figure 1

Figure 1
<p>Examples of aspheric sets: one convex non-smooth set and two non-convex smooth sets.</p>
Full article ">
21 pages, 7985 KiB  
Review
Mechanical Behavior of Flexible Fiber Assemblies: Review and Future Perspectives
by Peng Wang, Jiawei Han, Siyuan Wang and Yu Guo
Materials 2024, 17(24), 6042; https://doi.org/10.3390/ma17246042 - 10 Dec 2024
Viewed by 565
Abstract
Flexible fibers, such as biomass particles and glass fibers, are critical raw materials in the energy and composites industries. Assemblies of the fibers show strong interlocking, non-Newtonian and compressible flows, intermittent avalanches, and high energy dissipation rates due to their elongation and flexibility. [...] Read more.
Flexible fibers, such as biomass particles and glass fibers, are critical raw materials in the energy and composites industries. Assemblies of the fibers show strong interlocking, non-Newtonian and compressible flows, intermittent avalanches, and high energy dissipation rates due to their elongation and flexibility. Conventional mechanical theories developed for regular granular materials, such as dry sands and pharmaceutical powders, are often unsuitable for modeling flexible fibers, which exhibit more complex mechanical behaviors. This article provides a comprehensive review of the current state of research on the mechanics of flexible fiber assemblies, focusing on their behavior under compression, shear flow, and gas–fiber two-phase flow processes. Finally, the paper discusses open issues and future directions, highlighting the need for advancements in granular theories to better accommodate the unique characteristics of flexible fibers, and suggesting potential strategies for improving their handling in industrial applications. Full article
(This article belongs to the Section Soft Matter)
Show Figures

Figure 1

Figure 1
<p>Characteristic parameters for jammed flexible fiber materials: (<b>a</b>) coordination number <span class="html-italic">Z</span> varying with solid volume fraction <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math> from [<a href="#B4-materials-17-06042" class="html-bibr">4</a>], (<b>b</b>) mean pressure <math display="inline"><semantics> <mrow> <mi>p</mi> </mrow> </semantics></math> varying with <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math>, (<b>c</b>) critical solid volume fraction <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ϕ</mi> </mrow> <mrow> <mi>J</mi> </mrow> </msub> </mrow> </semantics></math> varying with fiber aspect ratio <span class="html-italic">AR</span> from [<a href="#B6-materials-17-06042" class="html-bibr">6</a>], and (<b>d</b>) solid volume fractions <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math> for packed flexible and rigid fibers from [<a href="#B7-materials-17-06042" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>Bulk mechanical and micromechanical responses [<a href="#B8-materials-17-06042" class="html-bibr">8</a>]: (<b>a</b>) quasistatic stress–strain cycles, and (<b>b</b>) coordination number as a function of time for the initial plastic cycles and the quasistatic regime (highlighted in orange).</p>
Full article ">Figure 3
<p>(<b>a</b>) Resistance force <math display="inline"><semantics> <mrow> <mi>F</mi> </mrow> </semantics></math> as a function of indentation depth <math display="inline"><semantics> <mrow> <mi>z</mi> </mrow> </semantics></math>, for granular chain assemblies with various numbers <span class="html-italic">N</span> of beads per chain [<a href="#B16-materials-17-06042" class="html-bibr">16</a>]. (<b>b</b>) Phase diagrams of yielding, transitional, and hardening regimes based on the fiber aspect ratio, <span class="html-italic">AR</span>, and fiber–fiber friction coefficient <math display="inline"><semantics> <mrow> <mi>μ</mi> </mrow> </semantics></math> for confining pressure <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math>= 10 kPa [<a href="#B18-materials-17-06042" class="html-bibr">18</a>]. The circle symbols in figure are the simulation results.</p>
Full article ">Figure 4
<p>Fiber shear flow schematics: (<b>a</b>) elongated rods at plane shear, where the blue and red dots indicate the two end faces of the cylinder. [<a href="#B27-materials-17-06042" class="html-bibr">27</a>]; (<b>b</b>) heat transfer of flexible fibers in a rotating drum, with particles of the same color representing those at the same temperature [<a href="#B28-materials-17-06042" class="html-bibr">28</a>]; (<b>c</b>) flexible filamentous particles in a rotary dryer with flights [<a href="#B29-materials-17-06042" class="html-bibr">29</a>].</p>
Full article ">Figure 5
<p>Solid-phase stress results of rigid rod-like particle systems (cylindrical or glued spheres) with different aspect ratios and no friction [<a href="#B37-materials-17-06042" class="html-bibr">37</a>]. (<b>a</b>) Kinetic components of the normalized normal stresses, (<b>b</b>) collisional component of the normalized normal stresses, (<b>c</b>) total stress of the normalized normal stresses, (<b>d</b>) kinetic components of the shear stresses, (<b>e</b>) collisional component of the shear stresses, and (<b>f</b>) total stress of the shear stresses as a function of solid volume fraction <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The schematic diagrams and probability distributions for inclination angle <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> and azimuthal angle <math display="inline"><semantics> <mrow> <mi>β</mi> </mrow> </semantics></math>: (<b>a</b>) schematic diagram of inclination angle <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> and azimuthal angle <math display="inline"><semantics> <mrow> <mi>β</mi> </mrow> </semantics></math>, (<b>b</b>) probability distributions for inclination angle <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> and azimuthal angle <math display="inline"><semantics> <mrow> <mi>β</mi> </mrow> </semantics></math> for the fibers with aspect ratios <span class="html-italic">AR</span> = 0.1, 1, and 6 [<a href="#B27-materials-17-06042" class="html-bibr">27</a>]; (<b>c</b>) probability distributions for inclination angle <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> and azimuthal angle <math display="inline"><semantics> <mrow> <mi>β</mi> </mrow> </semantics></math> for fibers with different particle aspect ratios <span class="html-italic">AR</span> and different solid volume fractions <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math> [<a href="#B37-materials-17-06042" class="html-bibr">37</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Variation of normalized shear stress with solid volume fraction for particles with and without friction. The surrounding images show snapshots of force chains at the specified solid volume fractions [<a href="#B27-materials-17-06042" class="html-bibr">27</a>]. (<b>b</b>) Comparison of stresses with rough glued-sphere, smooth glued-sphere particles, and cylindrical particles [<a href="#B37-materials-17-06042" class="html-bibr">37</a>]. (<b>c</b>) Snapshots from shear flows of flexible fibers at different solid volume fractions with or without friction [<a href="#B38-materials-17-06042" class="html-bibr">38</a>]. (<b>d</b>) DEM and experimental results of angles of repose of the wet fibers [<a href="#B39-materials-17-06042" class="html-bibr">39</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) The apparent friction coefficient and (<b>b</b>) normalized shear stress as a function of solid volume fraction for rigid and flexible fibers without friction. The average mechanical energies, including (<b>c</b>) elastic potential energy (PE), (<b>d</b>) local kinetic energy (LKE), (<b>e</b>) global kinetic energy (GKE), and (<b>f</b>) total mechanical energy (TME), per fiber as a function of solid volume fraction in the shear flows of rigid and flexible fibers with and without friction. Panels (<b>c</b>–<b>g</b>) shard a common legend [<a href="#B38-materials-17-06042" class="html-bibr">38</a>]. (<b>g</b>) Snapshots of the fibers with different flexibility. (<b>h</b>) Solid volume fraction and (<b>i</b>) average steady-state shear stress as a function of bond bending modulus at a normal stress of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi>y</mi> <mi>y</mi> </mrow> </msub> <mo>=</mo> <mn>1735</mn> <mo> </mo> <mi mathvariant="normal">P</mi> <mi mathvariant="normal">a</mi> </mrow> </semantics></math> [<a href="#B40-materials-17-06042" class="html-bibr">40</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Normalized shear stress as a function of solid volume fraction for various particle aspect ratios. (<b>b</b>) Variation of order parameter <span class="html-italic">S</span> with the maximum dimensional ratio of (<math display="inline"><semantics> <mrow> <mi>L</mi> <mo>/</mo> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <mo> </mo> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">d</mi> <mo> </mo> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math>) for dense flows at the solid volume fraction of <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>. (<b>c</b>) Normalized shear stresses as a function of the maximum value of <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>/</mo> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <mo> </mo> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">d</mi> <mo> </mo> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math> at different solid volume fractions [<a href="#B27-materials-17-06042" class="html-bibr">27</a>]. (<b>d</b>) Numerical models of shear tests of S-shaped fibers, the gray particles represent stationary particles, while the yellow particles indicate moving particles. (<b>e</b>) Sketches of S-shaped, U-shaped, and Z-shaped fibers. (<b>f</b>) Yield shear stress versus normalized maximum Feret diameter of a fiber <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>D</mi> </mrow> <mrow> <mi>F</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> <mo>/</mo> <msub> <mrow> <mi>l</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> </mrow> </semantics></math> for various sample lengths [<a href="#B43-materials-17-06042" class="html-bibr">43</a>].</p>
Full article ">Figure 10
<p>Snapshots of fluidized rod-like particles [<a href="#B50-materials-17-06042" class="html-bibr">50</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Snapshot of segregation results in a binary fiber mixed bed where yellow and purple represent particles with different diameters (<b>b</b>) segregation phase diagram [<a href="#B54-materials-17-06042" class="html-bibr">54</a>].</p>
Full article ">Figure 12
<p>Fibers in the bend and vertical pipe section at time 0.5 s colored by particle velocity, under different bonding stiffnesses [<a href="#B55-materials-17-06042" class="html-bibr">55</a>].</p>
Full article ">Figure 13
<p>Illustration of a flexible ribbon chain model. (<b>a</b>) Schematic diagram of the bond potential of the ribbon shaped particle and (<b>b</b>) flexible ribbon chain [<a href="#B56-materials-17-06042" class="html-bibr">56</a>].</p>
Full article ">Figure 14
<p>Spatial distributions of fibers in the riser at the superficial gas velocities of (<b>a</b>) 5.52 m/s and (<b>b</b>) 7.84 m/s [<a href="#B59-materials-17-06042" class="html-bibr">59</a>].</p>
Full article ">Figure 15
<p>Snapshots of fiber clusters with various (<b>a</b>) fiber aspect ratios (AR) and (<b>b</b>) dimensionless bending moduli (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>/</mo> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>f</mi> </mrow> </msub> <msubsup> <mrow> <mi mathvariant="bold-italic">U</mi> </mrow> <mrow> <mi>g</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math>) in a riser flow with the superficial gas velocity <b><span class="html-italic">U</span></b><span class="html-italic"><sub>g</sub></span> = 10 m/s. Average solid volume fraction is <span class="html-italic">ϕ<sub>a</sub></span> = 0.0758. The fibers are colored differently for visualization purposes [<a href="#B62-materials-17-06042" class="html-bibr">62</a>].</p>
Full article ">
16 pages, 8772 KiB  
Article
The Influence of Exogenous Particles on the Behavior of Non-Newtonian Mucus Fluid
by Agata Penconek, Urszula Michalczuk, Małgorzata Magnuska and Arkadiusz Moskal
Processes 2024, 12(12), 2765; https://doi.org/10.3390/pr12122765 - 5 Dec 2024
Viewed by 421
Abstract
Every day, approximately 7 m3 of air flows through the lungs of an adult, which comes into contact with 80 m2 of the lung surface. This air contains both natural and anthropogenic particles, which can deposit on the surface of the [...] Read more.
Every day, approximately 7 m3 of air flows through the lungs of an adult, which comes into contact with 80 m2 of the lung surface. This air contains both natural and anthropogenic particles, which can deposit on the surface of the mucus lining the respiratory tract. The presence of particles in the mucus leads to changes in its rheology and, consequently, in its functions. Therefore, this research aimed to determine how a non-Newtonian fluid suspension will behave during flow, illustrating the movement of mucus during coughing. The model mucus was an aqueous solution of carboxymethylcellulose (CMC). The tested particles suspended in a non-Newtonian fluid were Arizona Fine Dust, diesel exhaust particles, polyethylene microparticles, and pine pollen. It was noticed that as the fluid viscosity increases, the number of Kelvin–Helmholtz instabilities increases. The fluid’s expansion angle at the output of the measuring cell decreased, and the values of parameters characterizing the aerosol generated at the outlet decrease for selected particles present in the fluid. The research shows that the deposition of particles from polluted air in the respiratory tract, although they do not enter the bloodstream, may affect the human body. Deposited particles can change the behavior of mucus, which may translate into its functions. Full article
(This article belongs to the Special Issue Technological Processes for Chemical and Related Industries)
Show Figures

Figure 1

Figure 1
<p>SEM photo of (<b>A</b>) AFD, (<b>B</b>) PiPo, (<b>C</b>) DEP.</p>
Full article ">Figure 2
<p>(<b>A</b>) Experimental set-up, (<b>B</b>) the measurement cell.</p>
Full article ">Figure 3
<p>The characteristics of airflow.</p>
Full article ">Figure 4
<p>(<b>A</b>) The waves/instability occurring during the flow (red arrows marked the instability); (<b>B</b>) The maximum angle of aerosol spread at the outlet from the measurement cell.</p>
Full article ">Figure 5
<p>The apparent viscosity as a function of shear rate for 3.5%, 4.0%, and 4.5% CMC pure and with exogenous particles.</p>
Full article ">Figure 6
<p>The loss and storage modulus of 3.5% CMC pure and with exogenous particles.</p>
Full article ">Figure 7
<p>The loss and storage modulus of 4.0% CMC pure and with exogenous particles.</p>
Full article ">Figure 8
<p>The loss and storage modulus of 4.5% CMC pure and with exogenous particles.</p>
Full article ">Figure 9
<p>The number of instabilities compared to pure CMC.</p>
Full article ">Figure 10
<p>The angle of fluid ejection from the measuring cell.</p>
Full article ">Figure 11
<p>The droplet size distribution—3.5% CMC with exogenous particles.</p>
Full article ">Figure 12
<p>The droplet size distribution—4.0% CMC with exogenous particles.</p>
Full article ">Figure 13
<p>The droplet size distribution—4.5% CMC with exogenous particles.</p>
Full article ">Figure 14
<p>The dv10, dv50, d32 of aerosol at the outlet of measuring cell.</p>
Full article ">Figure 15
<p>(<b>A</b>) Sauter mean diameter (d32) vs. particle diameter; (<b>B</b>) Apparent viscosity vs. particle diameter.</p>
Full article ">Figure 16
<p>(<b>A</b>) Number of instabilities vs. apparent viscosity; (<b>B</b>) Aerosol spread angle vs. apparent viscosity.</p>
Full article ">
32 pages, 12061 KiB  
Article
Design of Trabecular Bone Mimicking Voronoi Lattice-Based Scaffolds and CFD Modelling of Non-Newtonian Power Law Blood Flow Behaviour
by Haja-Sherief N. Musthafa and Jason Walker
Computation 2024, 12(12), 241; https://doi.org/10.3390/computation12120241 - 5 Dec 2024
Viewed by 649
Abstract
Designing scaffolds similar to the structure of trabecular bone requires specialised algorithms. Existing scaffold designs for bone tissue engineering have repeated patterns that do not replicate the random stochastic porous structure of the internal architecture of bones. In this research, the Voronoi tessellation [...] Read more.
Designing scaffolds similar to the structure of trabecular bone requires specialised algorithms. Existing scaffold designs for bone tissue engineering have repeated patterns that do not replicate the random stochastic porous structure of the internal architecture of bones. In this research, the Voronoi tessellation method is applied to create random porous biomimetic structures. A volume mesh created from the shape of a Zygoma fracture acts as a boundary for the generation of random seed points by point spacing to create Voronoi cells and Voronoi diagrams. The Voronoi lattices were obtained by adding strut thickness to the Voronoi diagrams. Gradient Voronoi scaffolds of pore sizes (19.8 µm to 923 µm) similar to the structure of the trabecular bone were designed. A Finite Element Method-based computational fluid dynamics (CFD) simulation was performed on all designed Voronoi scaffolds to predict the pressure drops and permeability of non-Newtonian blood flow behaviour using the power law material model. The predicted permeability (0.33 × 10−9 m2 to 2.17 × 10−9 m2) values of the Voronoi scaffolds from the CFD simulation are comparable with the permeability of scaffolds and bone specimens from other research works. Full article
(This article belongs to the Section Computational Engineering)
Show Figures

Figure 1

Figure 1
<p>Trabecular bone has interconnected random porous structures of divergent pore sizes and different thicknesses of struts (trabeculae). Modified and reproduced with permission from Ref. [<a href="#B13-computation-12-00241" class="html-bibr">13</a>] CC BY 3.0.</p>
Full article ">Figure 2
<p>Zygoma and volumes of interest: Jugale (Ju), Middle Point (M.P.) and Zygomaxillare (Zm). Reproduced with permission from Ref. [<a href="#B14-computation-12-00241" class="html-bibr">14</a>] CC BY 4.0. The Ju area has the highest bone volume density (23.2 ± 4.3%) and highest trabecular plate thickness (0.16 ± 0.05 mm) in edentulous maxillae, compared to those at M.P. and Zm areas [<a href="#B15-computation-12-00241" class="html-bibr">15</a>,<a href="#B16-computation-12-00241" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>A Voronoi diagram and its tessellation patterns are based on random points to create the cells or regions in a given space. The boundary of each cell has an equal distance between two or more neighbouring points [<a href="#B40-computation-12-00241" class="html-bibr">40</a>].</p>
Full article ">Figure 4
<p>A Voronoi lattice-based cuboid biomimetic scaffold with functionally graded pores using a cloud of random seed points to create a stochastic microarchitecture.</p>
Full article ">Figure 5
<p>Zygoma model (front and back views) obtained from 3D scanning (EinScan-SE V2 3D scanner): Front view (<b>left</b>) and Back view (<b>right</b>) [<a href="#B14-computation-12-00241" class="html-bibr">14</a>].</p>
Full article ">Figure 6
<p>Virtual fracture process of zygoma: (<b>a</b>) fracture by a box, (<b>b</b>) fractured zygoma, and (<b>c</b>) extraction of the shape of the fracture for the design of Voronoi scaffolds (measurements in mm). (Refer: <a href="#app2-computation-12-00241" class="html-app">Appendix A</a> <a href="#computation-12-00241-f0A1" class="html-fig">Figure A1</a>).</p>
Full article ">Figure 7
<p>Random seed points inside the boundary layer with a distance known as point spacing. (Refer to <a href="#app3-computation-12-00241" class="html-app">Appendix B</a> <a href="#computation-12-00241-f0A2" class="html-fig">Figure A2</a>a).</p>
Full article ">Figure 8
<p>Conversion of scaffold shape (bone defect region) given on the left side into a volume mesh (Boundary layer) given on the right side. (Refer: <a href="#app3-computation-12-00241" class="html-app">Appendix B</a> <a href="#computation-12-00241-f0A2" class="html-fig">Figure A2</a>b).</p>
Full article ">Figure 9
<p>Creation of Voronoi lattices V90, V85, V80, V75 and V70 of porosities 90%, 85%, 80%, 75% and 70%, respectively. (Refer: <a href="#app3-computation-12-00241" class="html-app">Appendix B</a>).</p>
Full article ">Figure 10
<p>JT CAD files of different Voronoi scaffolds (front view) for CFD simulation.</p>
Full article ">Figure 11
<p>Surface wrapping method to create a fluid domain for the fluid surrounding the Voronoi V90 scaffold (solid domain).</p>
Full article ">Figure 12
<p>(<b>a</b>) Material properties of fluid and solid domains and (<b>b</b>) boundary conditions (velocity inlet, pressure outlet and no-slip wall condition) on the CFD model of the V90 scaffold.</p>
Full article ">Figure 13
<p>Generated volume mesh of the given (<b>a</b>) fluid domain and (<b>b</b>) solid domain of V90 scaffold for CFD simulation with tetrahedral elements and nodes.</p>
Full article ">Figure 14
<p>Scalar point maps and spheres are used to calculate pore sizes and pore numbers of all Voronoi lattice designs V90, V85, V80, V75, and V70. (Pore size calculation was carried out using the Lattice Pore size block of nTopology: Refer: <a href="#app3-computation-12-00241" class="html-app">Appendix B</a>).</p>
Full article ">Figure 15
<p>Graphs of variation of (<b>a</b>) surface area and (<b>b</b>) surface area/volume ratio (SA: V) with variation of porosities of Voronoi scaffolds.</p>
Full article ">Figure 16
<p>Testing of CFD models at different inlet velocities and the related pressure drops across the given Voronoi scaffolds.</p>
Full article ">Figure 17
<p>Velocity streamlines for all Voronoi scaffolds (V90, V85, V80, V75, V70) at inlet velocity 0.7 mm/s.</p>
Full article ">Figure 18
<p>Velocity streamlines for all Voronoi scaffolds (V90, V85, V80, V75, V70) at inlet velocity 0.7 mm/s.</p>
Full article ">Figure 19
<p>Pressure contours for all Voronoi scaffolds (V90, V85, V80, V75, V70) at inlet velocity of 0.7 mm/s.</p>
Full article ">Figure 20
<p>Variation of pressure drop and permeability based on the variation of porosity at an inlet velocity of 0.7 mm/s.</p>
Full article ">Figure 21
<p>Validation of a non-Newtonian model by a Newtonian model (comparison of pressure contours of both models for V90 scaffold) at inlet velocity of 0.7 mm/s.</p>
Full article ">Figure 22
<p>Shear strain rate contour of CFD model of V90 scaffold.</p>
Full article ">Figure A1
<p>(<b>a</b>) Creation of a virtual fracture of a scaffold using a box. (<b>b</b>) Fractured zygoma region. (<b>c</b>) Extracting the macrostructure or shape of the scaffold from bone defect or fractured region.</p>
Full article ">Figure A2
<p>(<b>a</b>) Creation of random seed points inside the implicit region. (<b>b</b>) Creation of boundary volume for Voronoi cells (Refer to <a href="#computation-12-00241-f0A3" class="html-fig">Figure A3</a>c). (<b>c</b>) Creation of Voronoi lattice in boundary volume adding strut thickness of 0.11 mm. (<b>d</b>) V70 Voronoi lattice.</p>
Full article ">Figure A3
<p>(<b>a</b>) Creation of pore diameters using scalar point maps. (<b>b</b>) Formation of spheres inside the Voronoi lattice, and (<b>c</b>) Voronoi cells inside the boundary volume.</p>
Full article ">Figure A4
<p>Calculation of porosity, volume fraction and SA: V.</p>
Full article ">
23 pages, 5577 KiB  
Article
The Influence of Temperature on Rheological Parameters and Energy Efficiency of Digestate in a Fermenter of an Agricultural Biogas Plant
by Maciej Filip Gruszczyński, Tomasz Kałuża, Wojciech Czekała, Paweł Zawadzki, Jakub Mazurkiewicz, Radosław Matz, Maciej Pawlak, Paweł Jarzembowski, Farokh Sahraei Nezhad and Jacek Dach
Energies 2024, 17(23), 6111; https://doi.org/10.3390/en17236111 - 4 Dec 2024
Viewed by 467
Abstract
This investigation specifically aims to enhance the understanding of digestate flow and mixing behavior across typical temperatures in bioreactors in agricultural biogas plants, facilitating energy-efficient mixing. Experimental tests confirmed that digestate exhibits non-Newtonian characteristics, allowing its flow behavior to be captured by rheological [...] Read more.
This investigation specifically aims to enhance the understanding of digestate flow and mixing behavior across typical temperatures in bioreactors in agricultural biogas plants, facilitating energy-efficient mixing. Experimental tests confirmed that digestate exhibits non-Newtonian characteristics, allowing its flow behavior to be captured by rheological models. This study validated that digestate rheology significantly varies with temperature, which influences flow resistance, mixing efficiency and overall energy requirements. Two rheological models—the Bingham and Ostwald models—were applied to characterize digestate behavior, with the Ostwald model emerging as the most effective for Computational Fluid Dynamic (CFD) simulations, given its balance between predictive accuracy and computational efficiency. Specifically, results suggest that, while three-parameter models, like the Herschel–Bulkley model, offer high precision, their computational intensity is less suitable for large-scale modeling where efficiency is paramount. The small increase in the accuracy of the shearing process description does not compensate for the significant increase in CFD calculation time. Higher temperatures were found to reduce flow resistance, which in turn enables increased flow rates and more extensive mixing zones. This enhanced mass transfer and mixing potential at elevated temperatures are especially pronounced in peripheral areas of the bioreactor, farthest from the agitators. By contributing a model for rheological behavior under realistic bioreactor conditions, this study supports the optimization of energy use in biogas production. These findings emphasize that temperature adjustments within bioreactors could serve as a reliable control strategy to maintain optimal production conditions while minimizing operational costs. Full article
(This article belongs to the Section A4: Bio-Energy)
Show Figures

Figure 1

Figure 1
<p>Przybroda agricultural biogas plant in Przybroda village, Poland.</p>
Full article ">Figure 2
<p>Comparison of the pseudo-curve and the real flow curve, and description of the real curve using the Ostwald power law model.</p>
Full article ">Figure 3
<p>Comparison of the course of the real flow curve and the description using the Newton, the Bingham and the Ostwald de Waele models.</p>
Full article ">Figure 4
<p>Comparison of the course of the real flow curve and description using the Bingham model.</p>
Full article ">Figure 5
<p>View of the computational domain of the tank used in numerical simulations of the digestate flow range.</p>
Full article ">Figure 6
<p>Area of measurement results for three parameters: shear stress, actual deformation gradient and temperature.</p>
Full article ">Figure 7
<p>The sample and circulating fluid temperature change as a function of time during the measurements. Measurements for two temperature settings with an upward difference of 2 °C.</p>
Full article ">Figure 8
<p>Graphs showing the fit of the models: linear (<b>a</b>,<b>c</b>) and exponential (<b>b</b>,<b>d</b>), describing the course of the rheological parameters of Bingham models: flow threshold and viscosity. Temperatures range from 30 °C to 56 °C.</p>
Full article ">Figure 9
<p>Area of fluid deformation results as a function of temperature of the Bingham model for the linear model. Temperatures range from 30 °C to 56 °C.</p>
Full article ">Figure 10
<p>Figures showing the fit of linear and exponential models: linear (<b>a</b>,<b>c</b>) and exponential (<b>b</b>,<b>d</b>), describing the course of rheological parameters of Ostwald models: consistency constant and flow rate. Temperatures range from 30 °C to 56 °C.</p>
Full article ">Figure 11
<p>Area of fluid deformation results as a function of temperature of the Ostwald model for the linear model. Temperatures range from 30 °C to 56 °C.</p>
Full article ">Figure 12
<p>Changes in fluid velocity profiles in the tank at distances of (<b>a</b>) 1 m, (<b>b</b>) 2 m, (<b>c</b>) 3 m and (<b>d</b>) 4 m from the inlet cross-section in the performed simulations.</p>
Full article ">Figure 13
<p>Changes in fluid flow velocity depending on the temperature in the tank at distances of (<b>a</b>) 3 m and (<b>b</b>) 4 m from the inlet cross-section in the performed simulations.</p>
Full article ">
27 pages, 36188 KiB  
Article
Back Analysis of a Real Debris Flow, the Morino-Rendinara Test Case (Italy), Using RAMMS Software
by Antonio Pasculli, Claudia Zito, Nicola Sciarra and Massimo Mangifesta
Land 2024, 13(12), 2078; https://doi.org/10.3390/land13122078 - 3 Dec 2024
Viewed by 478
Abstract
Debris flows are a dynamic and hazardous geological phenomenon, as by definition, debris flows are rapid, gravity-driven flows of saturated materials that often contain a mixture of water, rock, soil, and organic matter. They are highly destructive and occur in steep channels, posing [...] Read more.
Debris flows are a dynamic and hazardous geological phenomenon, as by definition, debris flows are rapid, gravity-driven flows of saturated materials that often contain a mixture of water, rock, soil, and organic matter. They are highly destructive and occur in steep channels, posing a significant threat to infrastructure and human life. The dynamics of debris flows are complex due to their non-Newtonian behaviour and varying sediment–water interactions, making accurate modelling essential for risk mitigation and emergency planning. This paper reports and discusses the results of numerical simulations of back analyses aimed at studying the reconstruction of a real rapid debris flow. The selected test case is the event that occurred on 12 and 16 March 2021 along the Rio Sonno channel, a tributary of the Liri River, following the landslide event of Rendinara (Municipality of Morino, Abruzzo Region, Italy). There are significant flow sources in the area, fed by a highly fractured carbonaceous aquifer that extends immediately upslope of the detachment zone. The continuous flow influences the saturation level in the fine-grained sediments and favours the triggering of the debris flow. This phenomenon was simulated using the commercial RAMMS code, and the rheological model selected was “Voellmy fluid friction”. The modelling approaches used in this research are valid tools to estimate the volumes of materials involved in the flow-feeding process and for the purpose of possible mitigation works (debris flow-type dams, weirs, flow diversion, etc.). Full article
(This article belongs to the Section Land – Observation and Monitoring)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The Roveto Valley located in the Abruzzo Region (Italy); (<b>b</b>) zoomed-in map of the landslide extension in the Rio Sonno/Fiume Liri area.</p>
Full article ">Figure 2
<p>Schematisation of tectonic lineament map.</p>
Full article ">Figure 3
<p>(<b>a</b>) Photo of the debris flow channel planimetry (Google Earth imagery); (<b>b</b>) longitudinal dimensions of the channel with the elevation.</p>
Full article ">Figure 4
<p>Sketch of the channel profile with the projected dimensions.</p>
Full article ">Figure 5
<p>Photos of some areas of the debris stream channel (Rio Sonno). (<b>a</b>) The upper part of the debris stream channel. (<b>b</b>) Zoomed-in photo of the selected area in <a href="#land-13-02078-f005" class="html-fig">Figure 5</a>a. (<b>c</b>) The lower part of the debris stream channel.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>): Bridge near the Liri River near the confluence of the Rio Sonno, along which the debris was channelled (red circle).</p>
Full article ">Figure 7
<p>Comparison of the SPH numerical solution, freely adapted from [<a href="#B56-land-13-02078" class="html-bibr">56</a>], and experimental results for a fast dam break, freely adapted from [<a href="#B81-land-13-02078" class="html-bibr">81</a>], in the time range from 0.1 s to 1 s, respectively, from (<b>a</b>) to (<b>b</b>).</p>
Full article ">Figure 8
<p>Discretisation map of the domain under study: detached (red) area; the discretised area included by the green lines, divided into two regions with different coefficients of friction; Region 1, area included by the dashed red lines; Region 2, area included by the dashed yellow lines.</p>
Full article ">Figure 9
<p>Assumed hydrograph, based on geophysical survey at the mouth between the Rio Sonno and Liri River.</p>
Full article ">Figure 10
<p>Flow height in the debris channel using a transitory time: 300 s (<b>a</b>), 400 s (<b>b</b>), 500 s (<b>c</b>), 600 s (<b>d</b>), 700 s (<b>e</b>) and 800 s (<b>f</b>).</p>
Full article ">Figure 11
<p>(<b>a</b>) Max. flow velocity; (<b>b</b>) max. flow pressure; (<b>c</b>) max. flow height, up to 800 s.</p>
Full article ">Figure 12
<p>Flow height at 600 s. Comparison between different material density equal to 1300 kg/m<sup>3</sup> (<b>a</b>) and material density equal to 1200 kg/m<sup>3</sup> (<b>b</b>).</p>
Full article ">Figure 13
<p>Flow height at 1000 s, density equal to 1300 kg/m<sup>3</sup>, and total debris mass equal to 100 m<sup>3</sup>.</p>
Full article ">Figure 14
<p>Comparison of the simulations between a velocity at 300 s (<b>a</b>) to a velocity at 1000 s (<b>b</b>).</p>
Full article ">Figure 15
<p>Flow height at 800 s, 1200 kg/m<sup>3</sup>; comparison between 200 m<sup>3</sup> (<b>a</b>) and 100 m<sup>3</sup> (<b>b</b>) as initial released debris volumes.</p>
Full article ">Figure 16
<p>Max. pressures (<b>a</b>) and max. velocities (<b>b</b>) after 800 s, assuming a density of 1200 kg/m<sup>3</sup> and a volume of 200 m<sup>3</sup> of released debris.</p>
Full article ">Figure 17
<p>Accumulation along the debris flow cross-section, shown in yellow in <a href="#land-13-02078-f017" class="html-fig">Figure 17</a>a. The grey fill is the cross-sectional flow height profile; the red line is the debris flow cross-sectional profile (multiple by 50) added to the debris channel topography; the green line is the cross-sectional debris channel topography. (<b>a</b>) Mouth of the debris flow. (<b>b</b>) Result for time 600 s and released mass 200 m<sup>3</sup>. (<b>c</b>) Result for time 700 s and released mass 100 m<sup>3</sup>. (<b>d</b>) Result for time 700 s and released mass 200 m<sup>3</sup>. (<b>e</b>) Result for time 800 s and released mass 100 m<sup>3</sup>. (<b>f</b>) Result for time 800 s and released mass 200 m<sup>3</sup>.</p>
Full article ">Figure 18
<p>Longitudinal height profile, along a selected line drawn along the Rio Sonno path. The red line is the debris flow longitudinal profile (multiple by 50) added to the debris channel topography; the green line is the debris channel topography. (<b>a</b>) Result for time 100 s and released mass 100 m<sup>3</sup>. (<b>b</b>) Result for time 100 s and released mass 200 m<sup>3</sup>. (<b>c</b>) Result for time 300 s and released mass 100 m<sup>3</sup>. (<b>d</b>) Result for time 300 s and released mass 200 m<sup>3</sup>. (<b>e</b>) Result for time 800 s and released mass 100 m<sup>3</sup>. (<b>f</b>) Result for time 800 s and released mass 200 m<sup>3</sup>.</p>
Full article ">
15 pages, 4530 KiB  
Article
Numerical Assessment of the Thermal Performance of Microchannels with Slip and Viscous Dissipation Effects
by Pamela Vocale and Gian Luca Morini
Micromachines 2024, 15(11), 1359; https://doi.org/10.3390/mi15111359 - 8 Nov 2024
Viewed by 661
Abstract
Microchannels are widely used across various industries, including pharmaceuticals and biochemistry, automotive and aerospace, energy production, and many others, although they were originally developed for the computing and electronics sectors. The performance of microchannels is strongly affected by factors such as rarefaction and [...] Read more.
Microchannels are widely used across various industries, including pharmaceuticals and biochemistry, automotive and aerospace, energy production, and many others, although they were originally developed for the computing and electronics sectors. The performance of microchannels is strongly affected by factors such as rarefaction and viscous dissipation. In the present paper, a numerical analysis of the performance of microchannels featuring rectangular, trapezoidal and double-trapezoidal cross-sections in the slip flow regime is presented. The fully developed laminar forced convection of a Newtonian fluid with constant properties is considered. The non-dimensional forms of governing equations are solved by setting slip velocity and uniform heat flux as boundary conditions. Model accuracy was established using the available scientific literature. The numerical results indicated that viscous dissipation effects led to a decrease in the average Nusselt number across all the microchannels examined in this study. The degree of reduction is influenced by the cross-section, aspect ratio and Knudsen number. The highest reductions in the average Nusselt number values were observed under continuum flow conditions for all the microchannels investigated. Full article
(This article belongs to the Section A:Physics)
Show Figures

Figure 1

Figure 1
<p>Schematic views of the parallel microchannels considered in the present analysis.</p>
Full article ">Figure 2
<p>Microchannels under investigation: (<b>a</b>) rectangular; (<b>b</b>) trapezoidal; (<b>c</b>) double trapezoidal.</p>
Full article ">Figure 3
<p>Nusselt number variation with number of mesh elements for rectangular microchannel with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0.01; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 4
<p>Nusselt number variation with number of mesh elements for trapezoidal microchannel with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0.01; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 5
<p>Nusselt number variation with number of mesh elements for double-trapezoidal microchannel with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0.01; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 6
<p>Comparison between the Poiseuille number (<span class="html-italic">f Re</span>) for square microchannels and the experimental data presented in [<a href="#B24-micromachines-15-01359" class="html-bibr">24</a>] for Nitrogen and Oxygen gases (i.e., <span class="html-italic">γ</span> = 1 and <span class="html-italic">Br</span> = 0).</p>
Full article ">Figure 7
<p>Variation of the dimensionless wall temperature for rectangular microchannels with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 8
<p>Variation of the dimensionless wall temperature for trapezoidal microchannels with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 9
<p>Variation of the dimensionless wall temperature for double-trapezoidal (or hexagonal) microchannels with <span class="html-italic">γ</span> = 0.1: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 10
<p><span class="html-italic">Nu</span> as a function of the aspect ratio in rectangular microchannels for different values of the Knudsen numbers: (<b>a</b>) <span class="html-italic">Br</span> = 0.01; (<b>b</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 11
<p><span class="html-italic">Nu</span> as a function of the aspect ratio in trapezoidal microchannels for different values of Knudsen numbers: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.01; (<b>c</b>) <span class="html-italic">Br</span> = 0.05; (<b>d</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 11 Cont.
<p><span class="html-italic">Nu</span> as a function of the aspect ratio in trapezoidal microchannels for different values of Knudsen numbers: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.01; (<b>c</b>) <span class="html-italic">Br</span> = 0.05; (<b>d</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 12
<p><span class="html-italic">Nu</span> as a function of the aspect ratio in double-trapezoidal (or hexagonal) microchannels for different values of the Knudsen numbers: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.01; (<b>c</b>) <span class="html-italic">Br</span> = 0.05; (<b>d</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">Figure 12 Cont.
<p><span class="html-italic">Nu</span> as a function of the aspect ratio in double-trapezoidal (or hexagonal) microchannels for different values of the Knudsen numbers: (<b>a</b>) <span class="html-italic">Br</span> = 0; (<b>b</b>) <span class="html-italic">Br</span> = 0.01; (<b>c</b>) <span class="html-italic">Br</span> = 0.05; (<b>d</b>) <span class="html-italic">Br</span> = 0.1.</p>
Full article ">
24 pages, 10140 KiB  
Article
On the Complex Flow Dynamics of Shear Thickening Fluids Entry Flows
by Miguel Montenegro and Francisco J. Galindo-Rosales
Micromachines 2024, 15(11), 1281; https://doi.org/10.3390/mi15111281 - 22 Oct 2024
Viewed by 819
Abstract
Due to their nature, using shear thickening fluids (STFs) in engineering applications has sparked an interest in developing energy-dissipating systems, such as damping devices or shock absorbers. The Rheinforce technology allows the design of customized energy dissipative composites by embedding microfluidic channels filled [...] Read more.
Due to their nature, using shear thickening fluids (STFs) in engineering applications has sparked an interest in developing energy-dissipating systems, such as damping devices or shock absorbers. The Rheinforce technology allows the design of customized energy dissipative composites by embedding microfluidic channels filled with STFs in a scaffold material. One of the reasons for using microfluidic channels is that their shape can be numerically optimized to control pressure drop (also known as rectifiers); thus, by controlling the pressure drop, it is possible to control the energy dissipated by the viscous effect. Upon impact, the fluid is forced to flow through the microchannel, experiencing the typical entry flow until it reaches the fully developed flow. It is well-known for Newtonian fluid that the entrance flow is responsible for a non-negligible percentage of the total pressure drop in the fluid; therefore, an analysis of the fluid flow at the entry region for STFs is of paramount importance for an accurate design of the Rheinforce composites. This analysis has been numerically performed before for shear-thickening fluids modeled by a power-law model; however, as this constitutive model represents a continuously growing viscosity between end-viscosity plateau values, it is not representative of the characteristic viscosity curve of shear-thickening fluids, which typically exhibit a three-region shape (thinning-thickening-thinning). For the first time, the influence of these three regions on the entry flow on an axisymmetric pipe is analyzed. Two-dimensional numerical simulations have been performed for four STFs consisting of four dispersions of fumed silica nanoparticles in polypropylene glycol varying concentrations (7.5–20 wt%). Full article
(This article belongs to the Special Issue Flows in Micro- and Nano-Systems)
Show Figures

Figure 1

Figure 1
<p>Sketch of the relationship between the velocity of the strike and the velocity of the fluid inside the microchannel.</p>
Full article ">Figure 2
<p>Steady shear viscosity curves for the four formulations as a function of shear rate. The data include different particle concentrations of (►) 7.5 wt%, (◇) 10 wt%, (■) 15 wt% and (○) 20%. Solid lines are fits of the high-rate-thinning model of Equation (5). Reprinted from [<a href="#B41-micromachines-15-01281" class="html-bibr">41</a>], with permission from Elsevier.</p>
Full article ">Figure 3
<p>Sketch of the channel geometry and boundary conditions.</p>
Full article ">Figure 4
<p>Mesh: (<b>a</b>) at the inlet region; (<b>b</b>) at the outlet region.</p>
Full article ">Figure 5
<p>Comparison of the mesh influence on normalized axial velocities and velocity profiles for STF 4 at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math> mm/s.</p>
Full article ">Figure 6
<p>Normalized axial velocity for the four formulations of four shear thickening fluids with increasing concentration for increasing inlet velocities (<b>a</b>) STF 1, (<b>b</b>) STF 2, (<b>c</b>) STF 3, and (<b>d</b>) STF 4.</p>
Full article ">Figure 7
<p>Shear-thickening region fitted to the power-law model (Equation (1)).</p>
Full article ">Figure 8
<p>Entrance length vs. Reynolds number.</p>
Full article ">Figure 9
<p>(<b>a</b>)—STF 4 viscosity along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity. (<b>b</b>)–STF 4 normalized shear stress along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity. (<b>c</b>)—STF 4 flow-type parameter along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>)—STF 4 viscosity along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity. (<b>b</b>)–STF 4 normalized shear stress along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity. (<b>c</b>)—STF 4 flow-type parameter along the pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm/s and matching normalized axial velocity.</p>
Full article ">Figure 10
<p>Evolution of peak normalized velocity with inlet velocity.</p>
Full article ">Figure 11
<p>Velocity profiles for the four formulations of shear thickening fluids with increasing concentration in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe: (<b>a</b>) STF 1, (<b>b</b>) STF 2, (<b>c</b>) STF 3, (<b>d</b>) STF 4.</p>
Full article ">Figure 12
<p>STF 1 normalized shear rate profile: (<b>a</b>) in the fully developed region for all inlet velocities and (<b>b</b>) at different <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math> mm/s.</p>
Full article ">Figure 13
<p>STF 1 normalized shear rate profiles at normalized coordinate <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>z</mi> </mrow> <mo>/</mo> <mrow> <mi>D</mi> </mrow> </mrow> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.2</mn> </mrow> </semantics></math> mm pipe for <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> mm/s. I, II and III represent each of the zones of the viscosity curve for a CST, according to Galindo-Rosales, et al. [<a href="#B40-micromachines-15-01281" class="html-bibr">40</a>,<a href="#B42-micromachines-15-01281" class="html-bibr">42</a>].</p>
Full article ">Figure 14
<p>Non-dimensional viscosity versus normalized shear rate along the wall for the formulations: (<b>a</b>) STF 1, (<b>b</b>) STF 2, (<b>c</b>) STF 3, and (<b>d</b>) STF 4.</p>
Full article ">Figure 15
<p>Loss coefficient at coordinate <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Normalized pressure drop evolution in the entry region (<math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>z</mi> <mo>≤</mo> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math>) for the four formulations of four shear thickening fluids with increasing concentration for increasing inlet velocities (<b>a</b>) STF 1, (<b>b</b>) STF 2, (<b>c</b>) STF 3, and (<b>d</b>) STF 4.</p>
Full article ">Figure 17
<p>Percentage of dissipated power along the tube in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe for the formulations: (<b>a</b>) STF 1; (<b>b</b>) STF 2; (<b>c</b>) STF 3; (<b>d</b>) STF 4.</p>
Full article ">Figure 18
<p>Percentage of dissipated power at coordinate <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <msub> <mrow> <mi>L</mi> </mrow> <mrow> <mi>e</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure A1
<p>Velocity profiles for the STF 1 formulation in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> (<b>a</b>) 0.1 mm/s, (<b>b</b>) 0.2 mm/s, (<b>c</b>) 0.3 mm/s, (<b>d</b>) 0.7 mm/s, (<b>e</b>) 1 mm/s, and (<b>f</b>) 1.5 mm/s.</p>
Full article ">Figure A2
<p>Velocity profiles for the STF 2 formulation in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> (<b>a</b>) 0.1 mm/s, (<b>b</b>) 0.2 mm/s, (<b>c</b>) 0.25 mm/s, and (<b>d</b>) 0.3 mm/s.</p>
Full article ">Figure A3
<p>Velocity profiles for the STF 3 formulation in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> (<b>a</b>) 0.1 mm/s, (<b>b</b>) 0.2 mm/s, (<b>c</b>) 0.25 mm/s, and (<b>d</b>) 0.3 mm/s.</p>
Full article ">Figure A4
<p>Velocity profiles for the STF 4 formulation in a <math display="inline"><semantics> <mrow> <mi>D</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> mm pipe: <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>v</mi> </mrow> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> (<b>a</b>) 0.1 mm/s, (<b>b</b>) 0.15 mm/s, (<b>c</b>) 0.2 mm/s, and (<b>d</b>) 0.25 mm/s.</p>
Full article ">
18 pages, 7115 KiB  
Article
The Numerical Simulation of the Injection Filling of the Fluidity Probe Die with Pattern Waxes
by Viacheslav E. Bazhenov, Arseniy S. Ovsyannikov, Elena P. Kovyshkina, Andrey A. Stepashkin, Anna A. Nikitina, Andrey V. Koltygin, Vladimir D. Belov and Dmitry N. Dmitriev
J. Manuf. Mater. Process. 2024, 8(5), 213; https://doi.org/10.3390/jmmp8050213 - 27 Sep 2024
Viewed by 1129
Abstract
Investment casting is a widely utilized casting technique that offers superior dimensional accuracy and surface quality. In this method, the wax patterns are employed in the layer-by-layer formation of a shell mold. As is customary, the patterns were created through the injection of [...] Read more.
Investment casting is a widely utilized casting technique that offers superior dimensional accuracy and surface quality. In this method, the wax patterns are employed in the layer-by-layer formation of a shell mold. As is customary, the patterns were created through the injection of molten or semi-solid wax into the die. The quality of the final casting is affected by the quality of the wax pattern. Furthermore, the filling of the die with wax can be associated with die-filling challenges, such as the formation of weld lines and misruns. In this study, the injection filling of the fluidity probe die with RG20, S1235, and S1135 pattern waxes was simulated using ProCast software. The thermal properties of the waxes, including thermal conductivity, heat capacity, and density across a wide temperature range, were determined with the assistance of a laser flash analyzer, a differential scanning calorimeter, and a dynamic mechanical analyzer. A favorable comparison of the acquired properties with those reported in the literature was observed. The Carreau model, which corresponds to non-Newtonian flow, was employed, and the parameters in the Carreau viscosity equation were determined as functions of temperature. Utilizing the thermal data associated with the wax patterns and the simulation outcomes, the interfacial heat transfer coefficients between the wax and the die were ascertained, yielding a value of 275–475 W/m2K. A strong correlation was observed between the experimental and simulated filling percentages of the fluidity probe across a wide range of injection temperatures and pressures. The analysis of the simulated temperature, fraction solid, viscosity, and shear rate in the wax pattern revealed that viscosity is a crucial factor influencing the wax fluidity. It was demonstrated that waxes with an initial high viscosity exhibit a low shear rate, which subsequently increases the viscosity, thereby hindering the wax flow. Full article
Show Figures

Figure 1

Figure 1
<p>The scheme of die utilized for the fluidity test: (<b>a</b>) the top and cross-sectional views, and (<b>b</b>) a three-dimensional representation of the probe and die.</p>
Full article ">Figure 2
<p>Digital vacuum wax injector (<b>a</b>) front view, (<b>b</b>) view without top showing wax pot configuration, and (<b>c</b>) mesh used for wax-filling simulation with set boundary conditions.</p>
Full article ">Figure 3
<p>(<b>a</b>) The exported frame with a vector image of the mold cavity after scaling; (<b>b</b>) experimental probe profile after tracing and Boolean intersection operation used for filled fraction calculation, and (<b>c</b>) the simulated wax fluidity probe filling results. All data (experimental and simulated) correspond to 1 s of filling the probe of RG20 wax at injection pressure and temperature of 196 kPa and 80 °C, respectively.</p>
Full article ">Figure 4
<p>Thermal properties of the die: (<b>a</b>–<b>c</b>) 5251 aluminum alloy calculated via ProCast thermodynamic database and (<b>d</b>–<b>f</b>) acrylic glass properties from references: Ameri (2021) [<a href="#B23-jmmp-08-00213" class="html-bibr">23</a>], McCabe (2008) [<a href="#B24-jmmp-08-00213" class="html-bibr">24</a>], and Al Sarheed (2022) [<a href="#B25-jmmp-08-00213" class="html-bibr">25</a>]. (<b>a</b>,<b>d</b>) density; (<b>b</b>) enthalpy; (<b>c</b>,<b>f</b>) thermal conductivity; (<b>e</b>) heat capacity.</p>
Full article ">Figure 5
<p>Thermal properties of the waxes obtained in this work (strength lines) and provided in the from references: Gebelin (2004) [<a href="#B12-jmmp-08-00213" class="html-bibr">12</a>], Gebelin (2003) [<a href="#B13-jmmp-08-00213" class="html-bibr">13</a>], Wang (2012) [<a href="#B14-jmmp-08-00213" class="html-bibr">14</a>], Bonilla (2001) [<a href="#B27-jmmp-08-00213" class="html-bibr">27</a>], Sabau (2003) [<a href="#B28-jmmp-08-00213" class="html-bibr">28</a>], Sabau (2002) [<a href="#B29-jmmp-08-00213" class="html-bibr">29</a>], Burlaga (2022) [<a href="#B30-jmmp-08-00213" class="html-bibr">30</a>], Ukrainczyk (2010) [<a href="#B31-jmmp-08-00213" class="html-bibr">31</a>] (dashed lines). (<b>a</b>) thermal conductivity; (<b>b</b>) heat capacity; (<b>c</b>)density.</p>
Full article ">Figure 6
<p>Influence of temperature on the parameters <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mo>∞</mo> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math>, <span class="html-italic">n</span>, λ in Carrea viscosity equation fitted by experimental viscosity data form [<a href="#B16-jmmp-08-00213" class="html-bibr">16</a>] for (<b>a</b>) RG20, (<b>b</b>) S1235, and (<b>c</b>) S1135 waxes.</p>
Full article ">Figure 7
<p>The fitted Carrea viscosity curves and experimental viscosity data form [<a href="#B16-jmmp-08-00213" class="html-bibr">16</a>] for (<b>a</b>) RG20, (<b>b</b>) S1235, and (<b>c</b>) S1135 waxes at various temperatures.</p>
Full article ">Figure 8
<p>The experimental and simulated cooling curves obtained for (<b>a</b>) RG20, (<b>b</b>) S1235, and (<b>c</b>) S1135 waxes. The simulated cooling curves were obtained at IHTC, determined by comparison of experimental and simulated temperatures.</p>
Full article ">Figure 9
<p>Experimental and simulated fluidity (filled fraction of the probe) of (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) RG20, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) S1235, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) S1135 waxes, depending on the time after filling start and injection temperatures at injection pressures. (<b>a</b>–<b>c</b>) P = 49 kPa, (<b>d</b>–<b>f</b>) P = 98 kPa, (<b>g</b>–<b>i</b>) P = 147 kPa, and (<b>j</b>–<b>l</b>) P = 196 kPa.</p>
Full article ">Figure 10
<p>Simulated (<b>a</b>–<b>d</b>) temperature, (<b>e</b>–<b>h</b>) fraction solid, (<b>i</b>–<b>l</b>) viscosity, (<b>m</b>–<b>p</b>) shear rate for (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>,<b>m</b>,<b>o</b>) RG20, and (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>,<b>n</b>,<b>p</b>) S1235 waxes for filling time (<b>a</b>,<b>b</b>,<b>e</b>,<b>f</b>,<b>i</b>,<b>j</b>,<b>m</b>,<b>n</b>) 1 s and (<b>c</b>,<b>d</b>,<b>g</b>,<b>h</b>,<b>k</b>,<b>l</b>,<b>o</b>,<b>p</b>) 2 s. The injection temperature and pressure were 90 °C and 196 kPa, respectively.</p>
Full article ">Figure 10 Cont.
<p>Simulated (<b>a</b>–<b>d</b>) temperature, (<b>e</b>–<b>h</b>) fraction solid, (<b>i</b>–<b>l</b>) viscosity, (<b>m</b>–<b>p</b>) shear rate for (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>,<b>m</b>,<b>o</b>) RG20, and (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>,<b>n</b>,<b>p</b>) S1235 waxes for filling time (<b>a</b>,<b>b</b>,<b>e</b>,<b>f</b>,<b>i</b>,<b>j</b>,<b>m</b>,<b>n</b>) 1 s and (<b>c</b>,<b>d</b>,<b>g</b>,<b>h</b>,<b>k</b>,<b>l</b>,<b>o</b>,<b>p</b>) 2 s. The injection temperature and pressure were 90 °C and 196 kPa, respectively.</p>
Full article ">
18 pages, 2268 KiB  
Article
Near Real-Time Estimation of Blood Loss and Flow–Pressure Redistribution during Unilateral Nephrectomy
by James Cowley, Justicia Kyeremeh, Grant D. Stewart, Xichun Luo, Wenmiao Shu and Asimina Kazakidi
Fluids 2024, 9(9), 214; https://doi.org/10.3390/fluids9090214 - 13 Sep 2024
Viewed by 735
Abstract
Radical or partial nephrectomy, commonly used for the treatment of kidney tumors, is a surgical procedure with a risk of high blood loss. The primary aim of this study is to quantify blood loss and elucidate the redistribution of blood flux and pressure [...] Read more.
Radical or partial nephrectomy, commonly used for the treatment of kidney tumors, is a surgical procedure with a risk of high blood loss. The primary aim of this study is to quantify blood loss and elucidate the redistribution of blood flux and pressure between the two kidneys and the abdominal aorta during renal resection. We have developed a robust research methodology that introduces a new lumped-parameter mathematical model, specifically focusing on the vasculature of both kidneys using a non-Newtonian Carreau fluid. This model, a first-order approximation, accounts for the variation in the total impedance of the vasculature when various vessels are severed in the diseased kidney (assumed to be the left in this work). The model offers near real-time estimations of the flow–pressure redistribution within the vascular network of the two kidneys and the downstream aorta for several radical or partial nephrectomy scenarios. Notably, our findings indicate that the downstream aorta receives an approximately 1.27 times higher percentage of the redistributed flow from the diseased kidney compared to that received by the healthy kidney, in nearly all examined cases. The implications of this study are significant, as they can inform the development of surgical protocols to minimize blood loss and can assist surgeons in evaluating the adequacy of the remaining kidney vasculature. Full article
(This article belongs to the Section Mathematical and Computational Fluid Mechanics)
Show Figures

Figure 1

Figure 1
<p>Schematic of the vascular network model assumed for the two kidneys and abdominal aorta. The left and right kidney were considered identical in structure and their major renal arteries Q<sub>R</sub> and Q<sub>L</sub> joined the aorta at a single location. For simplicity, the superior mesenteric artery was ignored. Each kidney was represented by an asymmetric vascular network composed of 25 vascular branching nodes (black circles) and 61 blood vessels (line segments) and organized into left (“L”) and right (“R”) branches relative to the first bifurcation node of each of the main renal arteries. The blood vessels of the left kidney were labelled as (LL1–LL23 and LR1–LR38), while those of the right kidney were labelled as (RL1–RL23 and RR1–RR38). See text for further details.</p>
Full article ">Figure 2
<p>(<b>A</b>) The vascular network for the human left kidney. The blue circle indicates an example of a bifurcation, where the blood vessel LR4 divides into the vessels LR6 and LR7, shown in more detail in (<b>B</b>) the healthy case and (<b>C</b>) a partial nephrectomy scenario where the branches LR6 and LR7 are cut. The presented first-approximation lumped-parameters model assumes that each vessel has an impedance described by (<b>B</b>) the three-element Windkessel model, with two resistors Z<sub>k1</sub> and Z<sub>k2</sub> and a compliance C<sub>k</sub> of the blood vessel. (<b>C</b>) The model also considers that each cut (severed) blood vessel has a zero output flux (modified from [<a href="#B53-fluids-09-00214" class="html-bibr">53</a>]).</p>
Full article ">Figure 3
<p>Schematic of the blood flux distribution for the healthy (uncut) two-kidney vascular network before any vessels in the left kidney are severed.</p>
Full article ">Figure 4
<p>Schematic of the blood flux redistribution due to severing a large portion of the outer regions of the left kidney (around the cortex and medulla), after the nodes Q<sub>LL6</sub>, Q<sub>LL9</sub>, Q<sub>LR2</sub>, and Q<sub>LR10</sub>, leaving intact the central region (the renal pelvis).</p>
Full article ">Figure 5
<p>Schematic of the blood flux redistribution between the unhealthy (<b>left</b>) kidney, the healthy (<b>right</b>) kidney, and the downstream aorta due to a radical left kidney resection with a cut after the node at Q<sub>L</sub>, the left major renal artery.</p>
Full article ">Figure 6
<p>Schematic of the blood pressure distributions for a healthy two-kidney vascular network, before any severing of blood vessels.</p>
Full article ">Figure 7
<p>Schematic of the blood pressure redistribution due to cross cutting after the nodes P<sub>LR2</sub>, P<sub>LR10</sub>, P<sub>LL6</sub>, and P<sub>LL9</sub> for the two-kidney vascular network.</p>
Full article ">Figure 8
<p>Schematic of the blood pressure redistribution due to a radical left renal resection with the severing of the major renal left artery after the node P<sub>L</sub>.</p>
Full article ">
Back to TopTop