Fact-checked by Grok 2 weeks ago

Characteristic length

In physics and , the characteristic length is a representative linear that defines the scale of a , serving as a key parameter in to form dimensionless groups for phenomena like fluid flow, , and mass transport. This length scale enables the prediction and scaling of system behavior without reliance on absolute sizes, facilitating comparisons between prototypes and full-scale models in fields such as and . In , the characteristic length is selected based on the of the —for instance, the for cylindrical or spheres, the chord length for airfoils, or the for non-circular ducts—and is to the (Re = \frac{\rho V L}{\mu}), where \rho is fluid density, V is , \mu is dynamic , and L is the characteristic length, to distinguish between laminar and turbulent flow regimes. Similarly, in applications, it is often defined as the ratio of volume to surface area (or area to perimeter for plates), as seen in natural correlations for the (Nu = \frac{h L}{k}), where h is the convective and k is , allowing of conduction dominance. The choice of characteristic length is context-dependent and must align with the dominant physical processes to ensure accurate similitude in experimental and computational modeling.

Core Concepts

Definition

A characteristic length, often denoted as L_c or L, is a single representative dimension that captures the essential size or scale of a or object, serving as a reference for laws, similarity , and the formation of dimensionless groups in and physics. It provides a standardized measure to compare phenomena across different geometries or conditions, enabling predictions of behavior without full-scale testing. For instance, it appears in dimensionless numbers like the to quantify the ratio of inertial to viscous forces in fluid flows. In applications, such as transient conduction problems, the characteristic length is often defined as L_c = \frac{V}{A}, where V is the volume of the body and A is the relevant surface area exposed to the process of interest. This volume-to-area ratio normalizes the scale of the system, yielding a that reflects how internal dimensions relate to boundary interactions, such as in conduction or problems. The choice of A depends on the physical context, ensuring the length aligns with the dominant transport mechanism. The application of characteristic length in was significantly advanced by in the early 20th century, where it played a key role in model testing and dynamic . Prandtl's 1904 presentation on boundary layers highlighted the need for such scales to simplify viscous flow analyses, bridging ideal fluid theory with real-world friction effects. Characteristic length is invariably expressed in units of length, such as (m) or feet (ft), to maintain dimensional consistency in equations. Common notations include L for general use, D when referring to a (e.g., in flows), or lowercase l in specific scaling contexts.

Role in

The characteristic length plays a pivotal role in by enabling the scaling of physical problems, allowing complex phenomena to be reduced to correlations through the formation of dimensionless parameters that remain independent of size. This approach facilitates the of results across different scales, as systems of varying sizes but with identical dimensionless groups exhibit similar behavior. For instance, in modeling, selecting an appropriate characteristic length ensures that predictive equations derived from small-scale experiments apply to larger prototypes without dimensional inconsistencies. In the framework of the Buckingham Pi theorem, the characteristic length serves as a fundamental repeating variable in the derivation of dimensionless pi groups, which systematically reduce the number of variables required to describe a physical relationship from an original set of n variables to n-m independent dimensionless combinations, where m is the number of fundamental dimensions. This theorem, formalized by Edgar Buckingham, underscores how incorporating a representative length scale—such as a system's diameter or height—allows for the elimination of dimensional dependencies, yielding relations solely among the pi groups. By choosing the characteristic length among the repeating variables, analysts can construct groups that capture essential interactions, such as inertial and viscous effects, thereby simplifying the analysis of multifaceted systems. The characteristic length further ensures similarity principles in experimental modeling, promoting geometric, kinematic, and dynamic similarity between scaled models and full prototypes. In applications like testing, scaling the characteristic length proportionally maintains matching dimensionless numbers, such as the , which governs flow regimes and force ratios across the model and actual system. This alignment allows experimental data to predict real-world performance reliably, as identical pi groups imply equivalent physical behaviors despite size differences. A primary example is the , where the characteristic length balances inertial and viscous forces to achieve dynamic similarity. However, the choice of characteristic length is not always unique, as multiple length scales may exist within a , potentially leading to inaccurate if an inappropriate one is selected. A poor selection can distort pi groups and compromise predictive accuracy, particularly in problems with disparate scales like . Despite these limitations, the use of characteristic length in promotes robust predictive modeling across scales by emphasizing scale-invariant correlations over absolute dimensions.

Selection and Determination

Principles for Choosing

The selection of a characteristic length in engineering analysis fundamentally depends on the dominant physical process governing the phenomenon, ensuring that it captures the scale over which key gradients—such as velocity, temperature, or concentration—occur most significantly. For instance, in convective heat transfer, the characteristic length often aligns with the direction of flow or the dimension perpendicular to heat flux boundaries, while in frictional flows, it relates to the wetted perimeter to represent shear effects. This choice must reflect the physics of the problem, such as inertia versus viscosity in momentum transport or conduction versus advection in energy transfer, to maintain physical relevance. Several factors influence the appropriate selection, including the complexity of the , the prevailing (laminar or turbulent), conditions, and the specific under consideration. Complex geometries may require a that averages local variations, such as the for non-circular ducts, defined as four times the cross-sectional area divided by the wetted perimeter, to approximate circular . In laminar flows, finer scales tied to layers may dominate, whereas turbulent regimes often favor larger geometric features like overall dimensions; conditions, such as no-slip walls, further dictate emphasis on near-wall scales. The in use, like the for inertial effects or the for convective enhancement, constrains the to ensure dimensional consistency. Common pitfalls in selection include arbitrary reliance on overall dimensions when local features, such as protrusions or thin layers, control the physics, leading to misrepresented gradients and inaccurate predictions. For example, using a room's as the for settling particles ignores the much smaller relevant to gravitational . To mitigate this, consistency must be maintained across interrelated parameters, such as pairing the length with a or that aligns with the same physical process. Theoretically, the characteristic length emerges from dimensional analysis, particularly the Buckingham π theorem, which groups variables into dimensionless forms that encapsulate essential physics without arbitrary units. This ensures the resulting dimensionless parameters, like the Reynolds or Prandtl numbers, accurately represent ratios of competing mechanisms, often validated through empirical correlations derived from experiments on geometries. Such selections enable scalable solutions applicable across similar systems, provided the underlying assumptions hold.

Geometry-Specific Methods

For simple geometric shapes, the characteristic length is typically selected as a representative linear that captures the of the object in the context of the physical process. For spheres and circular geometries, such as or cylinders in external , the D serves as the characteristic length, as it directly influences and in analyses. Similarly, for cubes and squares, the side length a is used, providing a consistent measure for bluff body interactions or across flat surfaces. In internal flows through non-circular ducts, the D_h is employed as the characteristic length to approximate the behavior of an equivalent circular duct, facilitating the use of standard correlations for and . The is defined as D_h = \frac{4A_c}{P}, where A_c is the cross-sectional area and P is the wetted perimeter. For a square duct with side length a, the cross-sectional area is A_c = a^2 and the wetted perimeter is P = 4a, yielding D_h = \frac{4a^2}{4a} = a. For a rectangular duct with dimensions a and b (where a > b), A_c = ab and P = 2(a + b), so D_h = \frac{4ab}{2(a + b)} = \frac{2ab}{a + b}. For more complex geometries where a single linear dimension is inadequate, the characteristic length can be derived from volumetric or surface properties to represent or scales. A common approach is the volume-to-surface area ratio, L_c = \frac{V}{A}, which quantifies the average distance from the interior to the boundary, particularly in heat conduction or porous media problems. For thin plates, where conduction dominates across the smallest dimension, the thickness t is selected as the characteristic length to assess gradients to the plate surface. In specialized applications like rocket combustion chambers, the characteristic length L^* = \frac{V_c}{A_t} is used, with V_c as the chamber volume and A_t as the throat area, to ensure complete mixing and efficiency. In computational approaches, such as finite element methods, the characteristic length guides refinement to balance accuracy and efficiency. For elements, it is often computed as L_c = V^{1/3}, where V is the element volume, providing a measure of local resolution; for elements, L_c = A^{1/2} with A as the area. This scaling helps control discretization errors in simulations of complex domains.

Applications in Disciplines

Fluid Dynamics

In fluid dynamics, the characteristic length L_c plays a pivotal role in the Reynolds number, a dimensionless parameter that characterizes the nature of fluid flow by balancing inertial and viscous forces. The Reynolds number is defined as Re = \frac{\rho v L_c}{\mu}, where \rho is the fluid density, v is the characteristic velocity, and \mu is the dynamic viscosity. This formulation arises from dimensional analysis applied to problems involving viscous drag, such as flow past a sphere or in a pipe. To derive it, consider the drag force F_d depending on \rho, \mu, v, and L_c; using the Buckingham Pi theorem with repeating variables \rho, v, and L_c, the dimensionless group emerges as \Pi = \frac{F_d}{\rho v^2 L_c^2} = f\left( \frac{\rho v L_c}{\mu} \right), yielding Re as the key parameter. Physically, Re represents the ratio of inertial forces (\rho v^2 L_c) to viscous forces (\mu v / L_c); low Re (typically below 2000–2300 for pipe flow) indicates laminar flow dominated by viscosity, while high Re (above 4000) signals turbulent flow where inertia prevails, with L_c setting the scale for these force interactions. The characteristic length also serves as the reference dimension in aerodynamic force coefficients, particularly for and . The is given by C_d = \frac{F_d}{\frac{1}{2} \rho v^2 A}, where A is the reference area, often scaled as L_c^2 (e.g., frontal area for bluff bodies or planform area for airfoils), ensuring C_d captures shape-dependent drag independently of size. Similarly, the coefficient C_l = \frac{F_l}{\frac{1}{2} \rho v^2 A} uses the same L_c-based area, allowing prediction of forces in scaled geometries. These coefficients depend on Re, as L_c influences the regime and boundary effects, with experimental data showing C_d decreasing with increasing Re due to reduced viscous dominance. In boundary layer theory, the characteristic length governs the development of the thin viscous layer near solid surfaces, where velocity gradients are significant. The \delta scales as \delta \sim L_c / \sqrt{[Re](/page/Re)}, derived from of the Navier-Stokes equations in the high-[Re](/page/Re) limit, where viscous effects confine to a region of order Re^{-1/2} relative to L_c. This scaling, from Prandtl's foundational work, underscores the layer's growth along the surface, affecting skin friction and separation. In airfoil design, controlling \delta via L_c (e.g., chord length) optimizes lift-to-drag ratios by delaying separation; in , it determines wall shear and efficiency, with thicker layers at lower [Re](/page/Re) increasing frictional losses. For experimental validation, the characteristic length enables similitude in scaled model testing, ensuring dynamic similarity through matched Re. In ship hull testing, models (e.g., 1:25 scale) are towed in basins to replicate full-scale viscous drag by adjusting velocity and L_c to equalize Re = \rho v L_c / \mu, often alongside Froude number matching for waves, allowing accurate extrapolation of resistance coefficients despite scale differences.

Heat and Mass Transfer

In and , the characteristic length L_c plays a pivotal role in dimensionless groups that quantify convective processes relative to conduction or . The , defined as Nu = \frac{h L_c}{k}, where h is the convective and k is the thermal conductivity, represents the enhancement of due to over pure conduction across the same length scale. The choice of L_c—such as the for cylindrical geometries—directly impacts the accuracy of empirical correlations for Nu, ensuring that predicted rates align with experimental data for specific shapes. In convection, L_c is incorporated into the , Gr = \frac{g \beta \Delta T L_c^3}{\nu^2}, where g is , \beta is the coefficient, \Delta T is the difference, and \nu is ; this number scales the -driven relative to viscous forces. Combined with the Pr = \frac{\nu}{\alpha} (where \alpha is ), Gr and Pr form the Ra = Gr \cdot Pr, which governs the onset and of convection flows, with L_c determining the cubic scaling of buoyancy effects. This incorporation allows correlations like Nu = f(Ra) to predict in scenarios such as vertical plates or enclosures, where L_c is typically the or gap width. The analogy between heat and mass transfer extends this framework to diffusive processes, where the Sherwood number Sh = \frac{k_m L_c}{D} mirrors the Nusselt number; here, k_m is the mass transfer coefficient and D is the diffusion coefficient, quantifying convective mass flux relative to molecular diffusion over L_c. This analogy holds for applications like drying, dissolution, or gas absorption, where correlations for Sh are derived similarly to Nu using Reynolds, Schmidt (analogous to Prandtl), and Grashof numbers, with L_c chosen based on geometry (e.g., particle diameter in packed beds). The Reynolds number, incorporating L_c, briefly influences the transition to turbulent convection in these mixed regimes. In heat exchanger design, selecting an appropriate L_c optimizes efficiency \eta_f = \frac{\tanh(m L_c)}{m L_c} (where m relates to and material properties), balancing extended surface area with drop along the to maximize overall transfer rates. For finned-tube exchangers, L_c as the height or ensures accurate prediction of and heat duty, guiding compact designs in applications like automotive radiators.

Computational Mechanics and Other Fields

In computational mechanics, the characteristic length L_c plays a crucial role in finite element analysis (FEA), particularly in controlling mesh refinement to accurately capture concentrations around geometric discontinuities or defects. The element size h, often taken as L_c, determines the resolution needed for reliable predictions, with finer meshes (smaller h) required near high-gradient regions to minimize discretization errors. In adaptive meshing strategies, L_c = h guides local refinement, ensuring that the concentration factor K_t—defined as the ratio of peak to nominal —is computed with sufficient accuracy without excessive computational cost. In within FEA frameworks, L_c addresses strain localization by regularizing mesh-dependent softening responses in damage models. A common approach scales the post-peak softening slope by L_c to preserve dissipation, where L_c may be the of the representative V^{1/3} for localized zones, preventing spurious sensitivity in simulations of quasi-brittle materials like . Seminal nonlocal models define L_c as the ratio of to the cohesive , typically on the order of millimeters for , enabling consistent predictions of damage evolution and crack propagation across scales. In rocket , the characteristic length L^* is a key parameter for designing combustion chambers in engines, defined as L^* = \frac{V_c}{A_t}, where V_c is the chamber volume and A_t the area. This length correlates with residence time, influencing efficiency and characteristic c^*, with optimal values around 1-1.5 meters for common bipropellants like / to achieve stable burning and high performance. Deviations from the ideal L^* can lead to incomplete or instability, as demonstrated in experimental studies optimizing lengths for bipropellant systems. Beyond these areas, characteristic lengths appear in acoustics, where the wavelength \lambda = \frac{c}{f} (with c as sound speed and f frequency) serves as L_c to characterize wave propagation and in enclosures or absorbers. In electromagnetics for design, L_c is typically a fraction of the wavelength, such as \lambda/2 for antennas, dictating and at operating frequencies. Emerging applications integrate L_c into , particularly (PINNs), where it scales input features and terms to normalize partial equations, improving for multi-scale problems like fluid-structure interactions. In PINNs for in composites, enhances estimation of L_c from microstructural , enabling accurate forecasting of without traditional FEA meshes. Studies show that varying L_c (e.g., or time scales) affects PINN accuracy, with adaptive mitigating errors in high-Reynolds flows.

Illustrative Examples

Internal Flows

In internal flows, such as those occurring in and channels, the characteristic length L_c is typically defined as the inner D for circular cross-sections, serving as the key in dimensionless for predicting . This choice of L_c = D enables the application of established correlations, including friction factor charts like the Moody diagram, which relate the to the and relative roughness for estimating s in systems. For instance, the \Delta P in a fully developed is calculated using the , \Delta P = f \frac{L}{D} \frac{\rho V^2}{2}, where f is the derived from charts that incorporate D as the scaling length, ensuring accurate predictions of energy losses over the pipe length L. For non-circular ducts and annuli, the D_h is employed as the characteristic length to generalize circular pipe correlations to irregular geometries, defined as D_h = \frac{4A}{P}, where A is the cross-sectional area and P is the wetted perimeter. This D_h allows the , \mathrm{Re} = \frac{\rho V D_h}{\mu}, to identify laminar or turbulent regimes in a manner consistent with circular pipes, facilitating the use of the same charts for calculations in rectangular ducts or annular spaces between concentric cylinders. In the specific case of between parallel plates separated by a gap width b, the characteristic length simplifies to L_c = 2b, equivalent to the , which accounts for the doubled effective dimension due to symmetric layers on both plates. In , the characteristically small L_c—often on the order of micrometers—amplifies surface-to-volume ratios, which profoundly influences flow dynamics by enhancing viscous and interfacial effects relative to inertial forces, typically resulting in laminar regimes even at moderate velocities. This leads to altered , such as rapid across the channel and dominance of wall stresses, necessitating specialized designs for applications like devices where the small L_c dictates mixing efficiency and heat transfer rates. A practical illustration of these concepts appears in HVAC duct design, where the characteristic length, often the of rectangular or circular ducts, directly impacts pressure losses and thus the required fan power to maintain specified rates. For example, increasing duct size to enlarge L_c reduces per unit , lowering the total drop and thereby decreasing fan by up to 20-30% in typical systems, as optimized designs balance with .

External Flows

In external flows, the characteristic length L_c plays a pivotal role in characterizing the interaction between a free stream and immersed objects, enabling the scaling of phenomena such as , , and through dimensionless numbers like the . Unlike confined internal flows, external configurations involve unbounded domains where the object's geometry dictates the dominant length scale, influencing development and . For blunt bodies like spheres and circular cylinders in crossflow, the characteristic length is typically the diameter D, serving as the reference for predicting drag coefficients via the Reynolds number Re_D = \frac{\rho U D}{\mu}. This choice arises because the diameter governs the scale of the adverse pressure gradient leading to flow separation, with wake formation and separation points directly scaling with D; for instance, the Strouhal number for vortex shedding frequency remains nearly constant over a range of Re when based on D. Experimental data confirm that drag crises, such as the drop in coefficient from approximately 1.2 to 0.1 for spheres at Re \approx 3 \times 10^5, are tied to boundary layer transition effects scaled by this length. In of airfoils and vehicles, the length c (the straight-line from leading to trailing ) is the characteristic length for calculations, forming the basis of the Re_c = \frac{\rho U c}{\mu} that determines sectional coefficients. This length influences angles by scaling the and separation bubble size; higher Re_c delays to larger angles of attack due to more stable laminar-to-turbulent transitions. For vehicle design, such as automobiles, the or frontal width analogously serves as L_c to correlate and with free-stream conditions. Atmospheric external flows around structures, like tall buildings, employ the building height H as the characteristic length for assessing wind loading, where Re_H = \frac{\rho U H}{\mu} helps predict mean and fluctuating pressures on facades. Gust scales at pedestrian levels, including turbulence intensity, diminish with distance from the structure but are fundamentally scaled by H, affecting comfort and load distribution in urban environments. A illustrative case is the in external underwater flows, where the fineness ratio \lambda = L/D (length to maximum ) modifies the effective characteristic length beyond a simple , optimizing minimization. For submerged operation, L_c is often taken as D for local effects, but the overall hull scales with \lambda \approx 6-8, as higher ratios reduce wave-making while viscous effects dominate at lower speeds; experimental models with \lambda = 11.3 demonstrate coefficients varying inversely with this ratio.

References

  1. [1]
    Fundamental Properties of Fluids – Introduction to Aerospace Flight ...
    However, the definition of a characteristic length scale may need to be qualified. Choosing a length scale, such as the flight vehicle's length, provides a ...
  2. [2]
    Dynamic Similarity – Introduction to Aerospace Flight Vehicles
    Remember that the Reynolds number is always based on some characteristic length (e.g., chord, mean chord, diameter, etc.) and should be qualified as such. Other ...
  3. [3]
    [PDF] Natural Convection Heat Transfer From an Isothermal Plate Aubrey ...
    The characteristic length L is the length scale of a physical system. For many heat transfer processes, it is the volume-to-surface-area or area-to-perimeter ...
  4. [4]
    Boundary Layer | Glenn Research Center - NASA
    Jul 17, 2024 · The theory which describes boundary layer effects was first presented by Ludwig Prandtl in the early 1900's. The general fluids equations had ...
  5. [5]
    Mach Number & Reynolds Number – Introduction to Aerospace ...
    Notice that when quoting the value of the Reynolds number, the characteristic length must be defined, e.g., the Reynolds number based on chord length, diameter, ...
  6. [6]
    None
    ### Definition of Characteristic Length in Heat Transfer
  7. [7]
    HISTORY OF BOUNDARY LA YER THEORY - Annual Reviews
    The boundary-layer theory began with Ludwig Prandtl's paper On the motion of a fluid with very small viscosity, which was presented at the Third International ...
  8. [8]
    [PDF] Ludwig Prandtl and Boundary Layers in Fluid Flow
    In 1904, Prandtl proposed the concept of bound- ary layers that revolutionised the study of fluid mechanics. In this article we present the basic ideas of ...
  9. [9]
    [PDF] Dimensional Analysis, Scaling, and Similarity - UC Davis Math
    The kinematic viscosity is a diffusivity of momentum; viscous effects lead to the diffusion of momentum in time T over a length scale of the order √ νT.
  10. [10]
  11. [11]
    [PDF] topic t3: dimensional analysis
    (3) Be able to carry out a formal dimensional analysis using Buckingham's Pi Theorem. ... characteristic length, L) and the fluid velocity, V, density ...
  12. [12]
    Similarity Parameters
    When wind tunnel testing, you must closely match the Mach number between the experiment and flight conditions. It is completely incorrect to measure a drag ...
  13. [13]
    [PDF] Fluid dynamics and heat transfer - Faculty members
    considering a constant density flow. Let's say the geometry of the ob- ject provides some characteristic length, `, and some characteristic flow velocity ...<|control11|><|separator|>
  14. [14]
    None
    Summary of each segment:
  15. [15]
    Chapter 4 Water flowing in pipes: energy losses
    For open channels the characteristic length is the hydraulic depth, the area of flow divided by the top width. For adequately turbulent conditions to exists, ...<|control11|><|separator|>
  16. [16]
    None
    Summary of each segment:
  17. [17]
    Viscous-Dominated Flows – Introduction to Aerospace Flight Vehicles
    For a sphere, the characteristic length scale L is taken as the sphere's diameter D , which is a key parameter in determining the flow regime around the body.
  18. [18]
    [PDF] Drag on Spheres - Penn State Mechanical Engineering
    Jan 11, 2012 · characteristic length of a body (for a sphere, L = d) m mass. Ma. Mach number: Ma = V/a. P static pressure. Patm atmospheric pressure. PT total ...
  19. [19]
    Worked Examples: Internal Flows – Introduction to Aerospace Flight ...
    Consult the Moody chart. (a) We are dealing with a duct that has a square cross-section, so we first need to find the hydraulic diameter D_h , which is.
  20. [20]
    [PDF] Thermally Developing Flow and Heat Transfer in Rectangular ...
    Feb 3, 2006 · rectangular channel and its hydraulic diameter are defined as b a α = and. 2 h ab. D. a b. = +. (1). Figure 1 shows a schematic diagram of the ...
  21. [21]
    [PDF] Home Experiment 2 Report Your Name ME 331: Introduction to Heat ...
    Apr 21, 2017 · Where Lc = V/A is a characteristic length. For a long cylinder the characteristic length is half the radius. In this case, the lumped ...
  22. [22]
    Characteristic Length and the Cooling Circle - Electronics Cooling
    Sep 11, 2012 · This means that heat behaves like a volume of water: A volume of ... heat the exisiting surface area and thus improve cooling. Cooling ...
  23. [23]
    [PDF] Liquid Engine Design: Effect of Chamber Dimensions on Specific ...
    L' is the length of the chamber. •Characteristic chamber length, L*, is the length that the chamber would be if it were a straight tube and had no converging ...
  24. [24]
    [PDF] CVEN 4511/5511: Introduction to Finite Element Analysis
    characteristic length of the mesh. Consider the 3D solid body with domain Ω, and discretize with 'elements' of characteristic length h (may not all be equal) ...
  25. [25]
    None
    ### Derivation of Reynolds Number from Dimensional Analysis for Viscous Flow
  26. [26]
    Transition and Turbulence - Princeton University
    In fluid flow, we often interpret the Reynolds number as the ratio of the inertia force (that is, the force given by mass x acceleration) to the viscous force.
  27. [27]
    The Drag Coefficient
    The drag coefficient is a number that aerodynamicists use to model all of the complex dependencies of drag on shape, inclination, and some flow conditions.
  28. [28]
    [PDF] Prandtl's Boundary Layer Theory - UC Davis Math
    Thus, the boundary layer thickness is of the order ε 1/2 , where is a length-scale ... Boundary layer theory applies in the limit Re -o . n that limit ...
  29. [29]
    Nusselt Correlations - Rose-Hulman
    The Nusselt number is defined as where h is the convective heat transfer coefficient, L chr is the appropriate characteristic length and k is the fluid thermal ...<|separator|>
  30. [30]
    [PDF] Page 1 Natural Convection Heat Transfer Net Force on a Body ...
    For a horizontal plate the characteristic length is δ=As/P where As and P are surface area and perimeter of the plate. ❑ Average Nusselt number correlations are.
  31. [31]
    The heat/mass transfer analogy factor, Nu/Sh, for boundary layers ...
    The equation states that local Nusselt numbers are proportional to local Sherwood numbers with ...
  32. [32]
    Fin Efficiency - an overview | ScienceDirect Topics
    An important consideration in the design of finned surfaces is selecting the proper fin length, because the fin efficiency decreases with the increase in fin ...
  33. [33]
    [PDF] HEAT EXCHANGER DIMENSIONING - Sistemas EEL
    where Af is the external surface area of the fin, and ηf the fin efficiency. Methods for calculating the fin efficiency can be found from any heat transfer ...
  34. [34]
    [PDF] A Technique for Mapping Characteristic Lengths to Preserve Energy ...
    At the subcell level, it is trivial to define the characteristic length, l. (βγ). C. —it is simply the length of the subcell perpendicular to the crack band.
  35. [35]
    [PDF] Estimating Stress Concentrations with a Minimal Mesh Density ...
    This tech brief outlines an approach for estimating stress concentrations without employing high mesh density finite element models. The method.
  36. [36]
    [PDF] MEASUREMENT OF CHARACTERISTIC LENGTH OF NONLOCAL ...
    The characteristic length I is the ratio of the fracture energy (i.e .• the energy dissipated per unit area. dimension N/m) to II.e energy dissipated per unit ...
  37. [37]
    (PDF) Characteristic Lengths of Liquid Propellant Rocket Engines ...
    Mar 31, 2023 · Characteristic length, L*, is a parameter used for design of combustion chambers of liquid propellant rocket engines. It is related to the stay ...
  38. [38]
    Optimum characteristic length of gas generator for liquid propellant ...
    The experimental results of hot tests show that the highest characteristic velocity of gas generator is achieved at a characteristic length of 4.27 m.
  39. [39]
    Determination of the viscous characteristic length in air‐filled porous ...
    Apr 1, 1996 · The concept of viscous characteristic length is used to describe the acoustical behavior of fluid‐saturated porous media in the high‐frequency ...
  40. [40]
    4.4: Antenna Characteristics - Engineering LibreTexts
    May 22, 2022 · An antenna designed to transmit and detect this light would need to be approximately of length λ 2 ≈ 250 ⁢ n ⁢ m . An atom is around ...
  41. [41]
    A machine learning enhanced characteristic length method for ...
    In this study, we demonstrate the benefit of machine learning methods to efficiently and accurately predict characteristic lengths of composite laminates with ...
  42. [42]
    The Apparent Pareto Front of Physics-Informed Neural Networks
    Aug 7, 2023 · In this paper, we aim to understand how parameters of the physical system, such as characteristic length and time scales, the computational ...
  43. [43]
    [PDF] Pipe Flows - Purdue Engineering
    Dec 15, 2021 · (3) For non-circular pipes or pipes ... Now consider a half-filled circular pipe of diameter D (Figure 11.11). The hydraulic diameter for.
  44. [44]
    [PDF] Fluid Mechanics MTF053 - Online Flow Calculator
    Hydraulic Diameter. For non-circular cross section ducts, the diameter D is replaced with the hydraulic diameter Dh calculated as. Dh = 4A. P. (1). The Reynolds ...<|separator|>
  45. [45]
    [PDF] a numerical investigation of turbulent flow in noncircular ducts
    Whereas the use of hydraulic diameter may cause some concern, these findings still suggest that a noncircular duct may actually be more efficient than circular ...
  46. [46]
    [PDF] Flow in pipe
    • For a developed flow between parallel plates: •. Streamlines are parallel ... Characteristic length = viscous length. $. %" e = Average wall roughness ...<|separator|>
  47. [47]
    [PDF] Microfluidics: Fluid physics at the nanoliter scale - Caltech
    Oct 6, 2005 · Each effect has its own characteristic time scale: ␶D=R0. 2/D for the beads to diffusively explore a length scale R0, ␶p=␰/kH for the ...
  48. [48]
    [PDF] MICROFLUIDICS
    low Re-numbers, i.e. viscous forces dominate, are typical for microfluidics. – l is a characteristic length scale. Page 29. Simplifications in microfluidics. 2.
  49. [49]
    [PDF] ASHRAEs-Duct-System-Design-Guide.pdf - HVAC Simplified
    Various aspects of the duct design process include minimizing fan energy use, as depicted in ... If the duct length between the fan outlet and the elbow is ...<|control11|><|separator|>
  50. [50]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    The aerodynamic behavior of a sphere is qualitatively similar to that of a circular cylinder, as shown in the figure. Again, there is a sudden drop in drag as ...
  51. [51]
    [PDF] 3ignature redacted - DSpace@MIT
    Feb 9, 2018 · For cylinders, the characteristic length is typically the diameter. Fig. 1-1 shows the drag coefficient of a smooth cylinder as a function of ...
  52. [52]
    [PDF] Computations of the drag coefficients for low-Reynolds-number flow ...
    In this section, the drag coefficients are computed for a sphere and a circular cylinder. The predictions are compared to previous experimental measurements and.
  53. [53]
    5. Chapter 5: Theory of Airfoil Lift Aerodynamics
    The chord length (c) is the length of the chordline and the common lengthscale for airfoils. Then the camberline is defined as the line through the mean ...
  54. [54]
    [PDF] 3,4. Low-Speed Aerodynamics, 2-D - Robert F. Stengel
    Unswept wing, symmetric airfoil, 2-D lift slope coefficient. Inviscid, incompressible flow. Referenced to chord length, c, rather than wing area. CL2−D = α.
  55. [55]
    [PDF] Large Eddy Simulation of Wind Loads on a Low-Rise Structure and ...
    where H denotes the characteristic length (e.g., the building height), f denotes the sampling frequency, and V denotes the mean wind velocity at the.
  56. [56]
    [PDF] Submarine Hull Design - NTNU
    Nov 27, 2015 · speed and L/D ≈ 6.5 for slow speed. For practical reasons (e.g. space requirements), a fineness ratio of L/D = 7 is scarcely met.
  57. [57]
    [PDF] Experimental Studies of the Drag of an Axisymmetric Submarine Hull
    Dec 13, 2002 · The model constructed for the current experiment was 286.47 mm (11.278 in) long with a diameter of 25.4 mm (1 in), which resulted in a fineness ...