Fact-checked by Grok 2 weeks ago

Reynolds number

The Reynolds number (Re) is a in that characterizes the behavior of fluid flow by representing the ratio of inertial forces to viscous forces, thereby predicting whether the flow regime is laminar, transitional, or turbulent. Introduced through experimental investigations into , it provides a critical scaling parameter for analyzing fluid motion across diverse scales and conditions, from microscopic channels to large-scale engineering systems. The concept originated from the work of Irish engineer and physicist Osborne Reynolds, who in 1883 presented a seminal paper to the detailing experiments with dyed flowing through glass to observe transitions between laminar and flow. Reynolds' apparatus demonstrated that the onset of depends on fluid velocity, pipe diameter, density, and , leading him to define a characteristic number that captures this dependency without units. Although the term "Reynolds number" was later formalized by in 1908, Reynolds' 1883 contribution remains foundational, influencing modern research and earning recognition in fields like and . Mathematically, the Reynolds number is expressed as Re = (ρ V L) / μ, where ρ is the fluid , V is a , L is a scale (such as pipe diameter), and μ is the dynamic viscosity of the fluid. An equivalent form uses kinematic viscosity ν = μ / ρ, yielding Re = V L / ν. Flow regimes are typically classified as laminar for Re < 2300, transitional for 2300 < Re < 4000, and turbulent for Re > 4000 in internal flows, though these thresholds can vary with and disturbances— for instance, turbulence in jets may be delayed until Re > 116,000. These values stem from Reynolds' original experiments and subsequent refinements, such as a 2011 study identifying the threshold for sustained in at approximately Re = 2040 (Avila et al.). The Reynolds number's significance lies in its ability to enable similarity in experiments and simulations, allowing engineers to model complex flows without full-scale prototypes; low Re values (e.g., below 100) dominate viscous effects in or biological systems like blood flow, while high Re (e.g., millions) governs inviscid approximations in design. Applications span for optimizing wing shapes and reduction, hydrodynamics in ship design, in exchangers, and like mixing in chemical reactors, where it informs friction factors, pressure drops, and . In modern contexts, it underpins (CFD) validations and , ensuring safe and efficient systems in , automotive, and .

Fundamentals

Definition

The Reynolds number, denoted as \operatorname{Re}, is a in that characterizes the nature of by representing the ratio of inertial forces to viscous forces within the fluid. This ratio helps determine whether the flow regime is laminar, transitional, or turbulent, providing a fundamental parameter for analyzing and predicting flow patterns. The standard mathematical formulation of the Reynolds number is given by \operatorname{Re} = \frac{\rho v D}{\mu}, where \rho is the fluid density, v is a characteristic velocity of the flow, D is a characteristic length scale (such as the diameter of a conduit), and \mu is the dynamic viscosity of the fluid. In this expression, the numerator \rho v D scales with inertial effects, while the denominator \mu scales with viscous effects, yielding a unitless value. Due to its dimensionless nature, the Reynolds number enables the scaling of flow behaviors across different physical sizes, velocities, and fluids, allowing engineers and scientists to predict similar hydrodynamic phenomena in models and prototypes without dependence on specific units. This universality underpins its widespread use in similitude analysis and in . The term "Reynolds number" derives from the name of Osborne Reynolds, the physicist and engineer who first identified the importance of this parameter in his seminal 1883 experimental study on flow transitions.

Derivation

The Reynolds number emerges as a dimensionless parameter in through several complementary derivation approaches, each rooted in scaling principles that highlight the balance between inertial and viscous forces.

Dimensional Analysis Approach

provides a systematic way to identify key dimensionless groups governing phenomena, without solving the governing equations explicitly. Consider a steady, characterized by \rho (dimensions [M L^{-3}]), a representative velocity v ([L T^{-1}]), a characteristic length scale D ([L]), and dynamic viscosity \mu ([M L^{-1} T^{-1}]). These four variables involve three dimensions (M, L, T), so the Buckingham \pi theorem predicts exactly one independent dimensionless group. To form this group, assume \pi = \rho^a v^b D^c \mu^d. Equating dimensions yields the system: \begin{align*} M: & \quad a + d = 0, \\ L: & \quad -3a + b + c - d = 0, \\ T: & \quad -b - d = 0. \end{align*} Solving gives a = 1, b = 1, c = 1, d = -1, so \pi = \rho v D / \mu, which is the Reynolds number \mathrm{[Re](/page/Re)}. This group encapsulates the ratio of inertial to viscous effects, serving as the primary scaling parameter for flow similarity.

Non-Dimensionalization of the Navier-Stokes Equations

An alternative derivation arises from scaling the Navier-Stokes equations, the fundamental governing equations for viscous fluid motion. The incompressible Navier-Stokes momentum equation is \rho \left( \frac{\partial \mathbf{u}}{\partial t} + \mathbf{u} \cdot \nabla \mathbf{u} \right) = -\nabla p + \mu \nabla^2 \mathbf{u}, where \mathbf{u} is , p is , and other symbols are as defined earlier. Introduce non-dimensional variables: \mathbf{x}^* = \mathbf{x}/L, \mathbf{u}^* = \mathbf{u}/U, t^* = t U / L, p^* = p / (\rho U^2), where L and U are and velocity scales. Substituting and multiplying through by L / (\rho U^2) yields the non-dimensional form: \frac{\partial \mathbf{u}^*}{\partial t^*} + \mathbf{u}^* \cdot \nabla^* \mathbf{u}^* = -\nabla^* p^* + \frac{1}{\mathrm{Re}} \nabla^{*2} \mathbf{u}^*, with \mathrm{Re} = \rho U L / \mu. Here, \mathrm{Re} multiplies the viscous diffusion term, directly quantifying the relative importance of inertial (order 1) to viscous forces. As \mathrm{Re} \to \infty, viscous effects diminish, approaching inviscid (Euler) flow; conversely, low \mathrm{Re} emphasizes .

Derivation via Momentum Balance in Boundary Layers

In boundary layer theory, the Reynolds number derives from order-of-magnitude balancing of terms in the simplified momentum equation near a solid surface, where viscous effects confine to a thin layer. The streamwise momentum equation approximates to u \frac{\partial u}{\partial x} + v \frac{\partial u}{\partial y} = -\frac{1}{\rho} \frac{\partial p}{\partial x} + \nu \frac{\partial^2 u}{\partial y^2}, neglecting streamwise diffusion and assuming slow transverse variations. Scale analysis uses x \sim L, u \sim U, y \sim \delta (), v \sim U \delta / L, with \nu = \mu / \rho. The inertial terms scale as U^2 / L, while the viscous term scales as \nu U / \delta^2. Balancing these gives U^2 / L \sim \nu U / \delta^2, so \delta / L \sim 1 / \sqrt{\mathrm{Re}}, where \mathrm{Re} = U L / \nu. This reveals \mathrm{Re} as the ratio of inertial to viscous stresses, determining the layer's relative thinness at high \mathrm{Re}. The pressure gradient term balances externally via the inviscid outer flow.

Choice of Characteristic Length

The characteristic length L in \mathrm{Re} must represent the scale over which flow gradients occur, ensuring consistent scaling. For circular pipes, L is the diameter D. For non-circular conduits, the hydraulic diameter D_h = 4A / P is used, where A is the cross-sectional area and P is the wetted perimeter; this preserves the balance in the viscous term for fully developed flow. For external flows, such as around objects, L is typically the body's dimension in the flow direction (e.g., chord length for airfoils). This choice ensures \mathrm{Re} captures geometry-dependent inertial-viscous interactions accurately.

History

The foundational understanding of laminar flow resistance in pipes emerged in the 1840s through independent experimental work by German hydraulic engineer Gotthilf Heinrich Ludwig Hagen and French physician Jean Léonard Marie Poiseuille. Hagen's 1839 investigations demonstrated that the flow rate of viscous fluids in narrow cylindrical tubes varied with the fourth power of the tube radius, establishing key empirical relations for steady, laminar motion without addressing transitions to turbulence. Poiseuille's contemporaneous studies from 1840 to 1841, motivated by blood flow in capillaries, similarly quantified pressure-driven laminar flow in small-diameter tubes, deriving proportional relationships between flow rate, pressure gradient, and tube dimensions that underscored viscous dominance in low-speed regimes. In 1883, British engineer and physicist Osborne Reynolds advanced this field by conducting pioneering experiments on the transition between laminar and turbulent flow in pipes. Using a horizontal glass tube through which water flowed under controlled velocity, Reynolds injected a thin stream of dye at the inlet to visualize flow patterns, observing that at low speeds the dye formed a straight, undisturbed filament indicative of laminar flow, while higher speeds caused sinuous distortions and eventual turbulent mixing. These experiments quantified a critical velocity threshold for the onset of turbulence, influenced by pipe diameter, fluid viscosity, and density, and were detailed in his seminal paper published first in the Proceedings of the Royal Society and expanded in the Philosophical Transactions. Reynolds' work earned recognition from the Royal Society, where he had been a Fellow since 1877, and it provided the empirical basis for a dimensionless parameter to characterize flow regimes. The dimensionless group derived from Reynolds' critical velocity—later formalized as the Reynolds number—evolved in nomenclature during the early among prominent fluid dynamicists. German physicist first termed it the "Reynolds number" in 1908 during a presentation on at the Congress of Mathematicians in , explicitly denoting the ratio as a pure central to stability analysis. , a leading figure in , adopted and popularized the term in his 1910 paper on analogies, referring to it as a established hydrodynamic parameter, and by 1913 explicitly attributed it to Reynolds while applying it to boundary layer flows and drag problems. This attribution solidified its naming and widespread use in theoretical fluid mechanics.

Flow Regimes in Conduits and Channels

Laminar–turbulent transition

The Reynolds number plays a central role in delineating flow regimes, where low values indicate dominance of viscous forces over inertial ones, sustaining orderly, with smooth streamlines. As the Reynolds number increases, inertial forces gain prominence, destabilizing the flow and promoting the onset of chaotic, turbulent motion characterized by eddies and mixing. This marks a fundamental shift in fluid behavior, with profound implications for , , and . The precise point of transition is quantified by critical Reynolds numbers, which serve as thresholds beyond which laminar flow becomes unstable. These values are not universal but depend on flow geometry, disturbance levels, and boundary conditions; for instance, in circular pipe flows, transition typically initiates around Re ≈ 2000 under controlled conditions and may extend to Re ≈ 4000 before fully turbulent flow establishes, reflecting the intermittent nature of the transitional regime. In boundary layers over flat plates, the critical Reynolds number based on momentum thickness is often around 520 for the onset of instability, though practical transitions occur at higher values due to environmental perturbations. Two primary mechanisms govern the transition process. In the classical pathway, small disturbances in the laminar shear layer evolve into Tollmien-Schlichting waves—viscous, streamwise-propagating instabilities predicted by theory—which amplify exponentially downstream, eventually nonlinearly distorting the flow and spawning turbulent spots. Alternatively, under elevated disturbance environments, bypass transition occurs, wherein intense external perturbations directly generate near-wall streaks and vortices, circumventing the orderly amplification of Tollmien-Schlichting waves and accelerating the breakdown to . These mechanisms, first theoretically analyzed by Tollmien in 1931 and Schlichting in 1933 for the wave instability, highlight the interplay between linear growth and nonlinear saturation in transition dynamics. Several factors modulate the critical Reynolds number and transition behavior. Surface roughness introduces localized disturbances that trip the , lowering the effective critical value and promoting earlier , particularly when roughness height exceeds a related to the local boundary-layer thickness. Entrance effects in confined flows, such as non-uniform inlet conditions in , generate initial perturbations that can shift the transition Reynolds number by up to 50% from ideal values, depending on inlet . Free-stream intensity further hastens transition by enhancing bypass mechanisms, with levels above 1% often reducing the critical Reynolds number substantially through direct momentum transfer into the . Geometry-induced gradients also influence , with adverse gradients accelerating and favorable ones delaying onset.

Flow in a pipe

In circular pipes, the Reynolds number determines the nature of the internal flow regime, influencing velocity profiles, pressure losses, and flow development. For low Reynolds numbers, typically < 2300, the flow is laminar, characterized by a parabolic velocity profile across the cross-section. This profile arises from the balance of viscous forces dominating over inertial effects, leading to smooth, streamlined layers of fluid moving parallel to the pipe axis. The Hagen-Poiseuille law governs this regime, deriving the axial velocity distribution as u(r) = \frac{\Delta P}{4 \mu L} (R^2 - r^2), where u(r) is the velocity at radial position r, \Delta P is the pressure drop over length L, \mu is the dynamic viscosity, and R is the pipe radius; this equation predicts the volumetric flow rate Q = \frac{\pi R^4 \Delta P}{8 \mu L}, emphasizing the strong dependence on pipe diameter to the fourth power. As the Reynolds number increases into the transitional range, approximately 2300 < Re < 4000, the flow becomes unstable, with intermittent bursts of turbulence disrupting the laminar structure, though the exact onset depends on entrance conditions and disturbances. Beyond this, for Re > 4000, the flow transitions to fully turbulent, exhibiting a flattened velocity profile near the pipe centerline due to enhanced momentum transfer from turbulent eddies, while a thin viscous sublayer persists near the wall. This profile results in higher centerline velocities relative to the mean, with the velocity defect law describing the outer region as \frac{u_{\max} - u}{u_{\tau}} = f\left( \frac{y u_{\tau}}{\nu} \right), where u_{\tau} is the friction velocity and y is the wall distance, though empirical correlations like the log-law are commonly used for practical predictions. The friction factor f, which quantifies wall in the Darcy-Weisbach \Delta P = f \frac{L}{D} \frac{\rho V^2}{2}, depends strongly on the Reynolds number and relative roughness \epsilon / D. For , f = 64 / \mathrm{Re}, a direct inverse relationship derived from Hagen-Poiseuille. In turbulent flow, f is determined from the , which compiles experimental data showing f decreasing with increasing Re for smooth pipes (following the Blasius correlation f \approx 0.316 / \mathrm{Re}^{0.25} for Re up to $10^5) and approaching a constant value in the fully rough regime at high Re, independent of Re. This chart enables engineers to predict pressure drops across a wide range of conditions by intersecting Re and \epsilon / D lines. The development of fully developed flow from the pipe entrance, known as the entrance length L_e, also scales with the Reynolds number. In laminar flow, L_e / D \approx 0.06 \mathrm{Re}, reflecting the gradual diffusion of vorticity from the wall to establish the parabolic profile over a distance proportional to Re. For turbulent flow, the shorter entrance length follows L_e / D \approx 4.4 \mathrm{Re}^{1/6}, as turbulent mixing accelerates profile development, typically resulting in L_e being only 10–60 diameters for common Re values.

Flow in a wide duct

In non-circular conduits such as rectangular ducts, the Reynolds number is defined using the D_h = 4A/P, where A is the cross-sectional area and P is the wetted perimeter, to characterize the regime effectively. This approach extends the standard Reynolds number formulation \mathrm{Re} = \rho V D_h / \mu, with \rho as fluid density, V as mean , and \mu as dynamic , allowing comparison to circular flows while accounting for geometric variations. For a rectangular duct with width a and height b, D_h = 2ab / (a + b), which approximates $2b for wide ducts where a \gg b. The transition from laminar to turbulent flow in wide rectangular ducts typically occurs around \mathrm{Re} \approx 2300 based on D_h, akin to circular pipes, but the duct's aspect ratio \alpha = a/b modifies the stability threshold. Experimental and numerical studies indicate that the critical Reynolds number decreases as the aspect ratio increases, with transition observed between \mathrm{Re} = 1765 and $2315 across aspect ratios from 1 to 10, due to sidewall-induced disturbances that promote earlier instability in wider configurations. For instance, in ducts with \alpha = 1 (square), transition aligns closely with the pipe value, while higher \alpha lowers the threshold, though inlet conditions and surface roughness can further influence the exact point. In fully developed laminar flow, the velocity profile resembles that in pipes but exhibits distinct corner effects in rectangular ducts, where the no-slip condition at all walls leads to a series solution of the Navier-Stokes equations, resulting in reduced velocities near corners and a more uniform core for wide aspect ratios. The axial velocity u(y,z) satisfies \nabla^2 u = (1/\mu) \mathrm{d}p/\mathrm{d}x, yielding a profile that is parabolic across the height but flattens along the width, deviating from the ideal parabolic form due to multi-wall interactions. In turbulent flow, the mean velocity follows a logarithmic law near walls, but non-square ducts develop secondary flows—counter-rotating vortices in the corners driven by anisotropy in the Reynolds stresses—which enhance momentum transfer and alter the primary profile, with vortex strength increasing for higher aspect ratios. These secondary motions, absent in fully developed laminar regimes, contribute up to 6% of the total wall shear stress in square ducts and become more pronounced in rectangular ones. Rectangular ducts are prevalent in heating, , and air-conditioning (HVAC) systems, where the Reynolds number based on D_h predicts flow regimes critical for and design, with wide ducts (\alpha > 5) often operating at \mathrm{Re} up to 500,000 in turbulent conditions. In such applications, influences friction factors, which increase monotonically with \alpha at fixed \mathrm{Re}, necessitating adjustments to standard correlations for accurate sizing.

Flow in an open channel

In open channel flow, which is gravity-driven with a free surface exposed to the atmosphere, the Reynolds number is defined using the hydraulic radius as the characteristic length scale to account for the irregular geometry of the cross-section. The hydraulic radius R_h is given by R_h = \frac{A}{P}, where A is the flow cross-sectional area and P is the wetted perimeter. The Reynolds number is then expressed as \mathrm{Re} = \frac{v R_h}{\nu}, with v denoting the mean flow velocity and \nu the kinematic viscosity of the fluid. Laminar flow in open channels is rare and generally occurs only when \mathrm{Re} < 500, whereas turbulent flow dominates for \mathrm{Re} > 2000, with the intermediate range representing transitional conditions. These thresholds are adapted from criteria but adjusted for the free surface, where the transition is further modulated by the \mathrm{Fr} = \frac{v}{\sqrt{g R_h}}, which captures the relative of inertial to gravitational forces and influences subcritical or supercritical regimes that interact with viscous effects. For instance, low \mathrm{Re} combined with low \mathrm{Fr} (<1) typically yields laminar-subcritical flow, while higher values lead to turbulent-subcritical conditions prevalent in natural streams. In turbulent open channel flows, velocity profiles vary with bed conditions. Over rough beds, the near-bed region exhibits a logarithmic profile, described by the law of the wall: \frac{u}{u_*} = \frac{1}{\kappa} \ln \left( \frac{y u_*}{\nu} \right) + B, where u is the local velocity at height y above the bed, u_* = \sqrt{\tau_0 / \rho} is the shear velocity (\tau_0 bed shear stress, \rho fluid density), \kappa \approx 0.4 is von Kármán's constant, and B is a roughness-dependent constant./04:_Flow_in_Channels/4.07:_Velocity_Profiles) In wide channels where sidewall effects are negligible, the profile becomes nearly uniform across most of the depth, with shear concentrated near the bed, approximating plug flow away from the boundary layer./04:_Flow_in_Channels/4.07:_Velocity_Profiles) Bed roughness and channel slope significantly influence Re-based regime classification by altering the mean velocity and shear distribution for a given discharge. Increased roughness height relative to R_h (high relative roughness k_s / R_h) shifts the flow toward the fully rough turbulent regime at lower Re, as viscous effects diminish and form drag dominates, per Nikuradse's experiments adapted to channels. Steeper slopes elevate v through the uniform flow relation (e.g., Manning's equation v = \frac{1}{n} R_h^{2/3} S^{1/2}, with S bed slope and n roughness coefficient), thereby increasing Re and favoring turbulence, though excessive slope can couple with Fr to induce supercritical transitions that amplify instability.

Interactions with Objects and Boundaries

Flow around airfoils

In the context of flow around airfoils, the Reynolds number is defined using the chord length c as the characteristic length, given by the formula \text{Re} = \frac{v c}{\nu}, where v is the freestream velocity and \nu is the kinematic viscosity of the fluid. This dimensionless parameter governs the balance between inertial and viscous forces along the airfoil surface, influencing boundary layer behavior and overall aerodynamic performance. For typical aircraft applications, Reynolds numbers range from approximately $10^5 for small unmanned aerial vehicles (UAVs) to $10^8 for large commercial jets, reflecting variations in scale, speed, and altitude. At low Reynolds numbers, typically below $10^6, the boundary layer remains predominantly laminar, leading to thicker profiles and increased susceptibility to separation. This often results in the formation of laminar separation bubbles near the leading edge, where the flow detaches, undergoes transition to turbulence, and may reattach, creating a region of recirculating flow that elevates drag coefficients significantly—sometimes by factors of 2–5 compared to higher-Re conditions. Such effects are pronounced in applications like drones and gliders operating at Re ≈ $10^5–$3 \times 10^5, where the reduced momentum in the boundary layer exacerbates adverse pressure gradients, causing earlier stall and limiting maximum lift-to-drag ratios to below 20 in some cases. In contrast, at high Reynolds numbers exceeding $10^6, the boundary layer transitions to turbulence earlier, enhancing its resistance to separation through increased mixing and momentum transfer near the wall. This turbulent boundary layer can withstand stronger adverse pressure gradients, delaying flow separation on the airfoil's aft section and enabling higher lift generation before stall occurs. For instance, in transonic flows over transport aircraft airfoils at Re ≈ $10^7–$10^8, the turbulent layer reduces shock-induced separation, improving pressure recovery and overall efficiency. Reynolds number mismatches between wind tunnel models and full-scale flight introduce significant scale effects, often overpredicting drag and underpredicting lift in sub-scale tests due to delayed transition and premature separation at lower model Re (e.g., $10^6 vs. $10^8). These discrepancies necessitate corrections, such as trip strips to force transition or pressurized tunnels to elevate Re, ensuring closer alignment with flight conditions for accurate aerodynamic predictions.

Object in a fluid

When an arbitrary object is immersed in a fluid flow, the (Re = ρ U L / μ, where ρ is fluid density, U is flow velocity, L is a characteristic length, and μ is dynamic viscosity) governs the balance between inertial and viscous forces, profoundly influencing the resulting hydrodynamic forces and wake structures. At low (Re ≪ 1), viscous forces dominate, leading to creeping or where streamlines are symmetric fore and aft of the object, with no flow separation and drag primarily due to skin friction. In this regime, the drag coefficient C_D (defined as C_D = F_D / (½ ρ U² A), with F_D the drag force and A the projected area) scales inversely with Re, typically C_D ≈ k / Re for some constant k depending on shape, as inertial effects are negligible. Lift coefficient C_L, similarly nondimensionalized, is generally zero for symmetric objects in this viscous-dominated flow due to fore-aft symmetry. As Re increases into transitional regimes (roughly 1 < Re < 10^5), inertial effects become significant, causing regime shifts where a forms around the object and separation may occur, leading to asymmetric wakes and elevated . In the inertial regime (Re ≫ 1, often >10^5 for bluff bodies), viscous effects are confined to thin s, and C_D approaches a nearly constant value (Newton's regime, C_D ≈ 0.4–1.2 depending on shape), dominated by (form) from separated rather than viscous shear. For objects with asymmetry or , C_L can become substantial and Re-dependent, as growth and separation alter distributions; however, at very high Re (>10^5), a may occur where transition to suddenly reduces C_D by delaying separation. These shifts highlight how Re dictates the transition from viscosity-controlled to inertia-controlled forces, with general classifications including the Stokes regime (Re < 1, creeping ), transitional regime (1 < Re < 10^5, vortex formation and shedding), and fully turbulent regime (Re > 10^5, chaotic wakes with turbulent s). Wake formation behind the object evolves distinctly with Re. In the Stokes regime, the wake is steady and symmetric with no recirculation. At intermediate Re (typically 10 < Re < 300 for many shapes), a steady pair of attached vortices forms in the wake due to early separation. Beyond this (e.g., Re > 40–50), periodic vortex shedding emerges, creating alternating vortices in a von Kármán vortex street that extends downstream. This shedding frequency f is characterized by the Strouhal number St = f L / U, which remains nearly constant at St ≈ 0.18–0.22 over several orders of magnitude in Re (10² to 10⁵) for bluff bodies, reflecting a universal inertial scaling in the wake instability. At higher Re (>10^5), the wake becomes fully turbulent, with irregular shedding and enhanced mixing. Boundary layer separation, the point where flow detaches from the object's surface, critically depends on both Re and object shape, determining wake size and force coefficients. For bluff (non-streamlined) shapes like cylinders or blocks, separation occurs early due to sharp adverse pressure gradients, even at moderate Re, leading to large wakes and high form drag. Streamlined shapes delay separation via gradual pressure recovery, but at low Re, thicker laminar boundary layers separate more readily; conversely, at high Re, transition to a turbulent boundary layer (thinner and more resistant to adverse gradients) allows flow attachment longer, reducing drag. This Re-shape interplay explains why drag and wake characteristics vary: for instance, angular bluff bodies exhibit Re-insensitive separation points and constant C_D, while smooth ones show stronger Re dependence through boundary layer transitions.

Sphere in a fluid

The flow of a past a is a canonical problem in , where the Reynolds number, defined as Re = \frac{2 \rho v r}{\mu} with \rho as density, v as , r as , and \mu as dynamic , governs the transition from viscous-dominated to inertia-dominated regimes. At low Reynolds numbers, inertial effects are negligible, leading to creeping flow where viscous forces balance the motion. For Re < 1, the drag force on the sphere is given by : F_d = 6 \pi \mu r v, derived from solving the under no-slip boundary conditions, providing an exact analytical solution for steady, incompressible flow. This law accurately predicts the drag in highly viscous fluids, such as the settling of fine particles in sedimentation experiments. At intermediate Reynolds numbers, approximately $1 < Re < 10, inertial effects begin to influence the wake behind the sphere, invalidating the pure . Oseen's approximation addresses this by linearizing the convective terms in the around the uniform upstream flow, yielding a corrected drag force of F_d = 6 \pi \mu r v \left(1 + \frac{3}{16} Re \right), which improves accuracy by accounting for inertia in the far field while retaining viscous dominance near the sphere. For higher Reynolds numbers, empirical relations are essential, as analytical solutions become infeasible. The drag coefficient C_d = \frac{F_d}{\frac{1}{2} \rho v^2 \pi r^2} versus Re follows a characteristic curve: at low Re, C_d \approx 24/Re, matching ; it then decreases gradually through transitional regimes before stabilizing around C_d \approx 0.4 for $10^3 < Re < 10^5, where a separated wake forms but remains axisymmetric. Beyond Re \approx 3 \times 10^5, a drag crisis occurs, causing C_d to drop sharply to about 0.1 due to boundary layer transition to turbulence, delaying separation, before rising again at even higher Re. In settling experiments, the terminal velocity v_t of a sphere is reached when the drag force balances the net gravitational force (\rho_p - \rho) \frac{4}{3} \pi r^3 g, with \rho_p as particle density and g as gravity. For low Re, yields v_t = \frac{2 r^2 (\rho_p - \rho) g}{9 \mu}, enabling direct measurement of viscosity from observed fall speeds in viscometers. At higher Re, iterative solutions using the empirical C_d(Re) curve are required to compute v_t, as demonstrated in particle settling studies across fluid regimes.

Rectangular object in a fluid

The flow of a fluid over a rectangular object, such as a prism or plate, is characterized by the Reynolds number (Re), typically defined using the object's characteristic width D perpendicular to the flow direction, the fluid velocity v, density \rho, and dynamic viscosity \mu, as \operatorname{Re} = \rho v D / \mu. At low Reynolds numbers, the flow remains laminar and adheres closely to the no-slip boundary condition at the object's surfaces, leading to significant viscous drag and a broad wake due to the dominance of inertial forces being suppressed. This results in high drag coefficients, often exceeding 2, as the viscous effects extend far downstream without significant separation bubbles forming at the edges. As the Reynolds number increases beyond approximately 40–200, the flow undergoes a transition where boundary layer separation occurs prominently at the sharp edges of the rectangular body, initiating unsteady vortex shedding and the formation of a in the wake. This periodic shedding arises from the instability of the separated shear layers, alternating from the leading and trailing edges, and marks a shift from steady to oscillatory flow patterns, with the wake becoming more turbulent at higher Re. The exact onset depends on the body's geometry but typically aligns with this range for bluff rectangular shapes, contrasting with smoother bodies where transitions occur at slightly lower values. The aspect ratio of the rectangular object—defined as the ratio of its depth (streamwise length) to width D—significantly influences the flow regime and Reynolds number scaling. For infinite (two-dimensional) cylinders, where end effects are negligible, Re is based solely on D, and the flow exhibits pronounced two-dimensional vortex shedding. In contrast, finite-width plates introduce three-dimensional end effects, reducing the effective Re influence and altering separation patterns, with drag moderated by spanwise flows; studies show that higher aspect ratios (longer depth relative to width) stabilize the wake and diminish vortex street intensity compared to squat prisms. In the intermediate Reynolds number regime of $10^3 to $10^5, the drag coefficient C_d for rectangular objects stabilizes at values approximately 1–2, reflecting a balance between form drag from the separated wake and skin friction, with minimal variation until supercritical transitions at higher Re. Vortex shedding frequency is quantified by the St ≈ 0.2, which governs the shedding rate via the relation f = \frac{\text{St} \, v}{D}, where f is the shedding frequency; this near-constant St holds across a range of aspect ratios for subcritical flows, enabling prediction of oscillatory loads. These Reynolds number-dependent phenomena are critical in engineering applications, particularly for assessing wind loading on bluff structures like buildings or bridges modeled as rectangular prisms, where low-Re viscous effects amplify base drag, while vortex-induced vibrations from shedding at higher Re necessitate design mitigations to prevent structural fatigue. Experimental data from high-pressure wind tunnels confirm that Re sensitivity persists up to $10^6, influencing mean and fluctuating forces on such bodies.

Fall velocity

The terminal velocity v_t of a particle settling in a fluid is reached when the net downward force due to gravity and buoyancy balances the upward drag force. For a spherical particle of radius r and density \rho_p in a fluid of density \rho and dynamic viscosity \mu, the gravitational force is \frac{4}{3}\pi r^3 \rho_p g, while the buoyant force is \frac{4}{3}\pi r^3 \rho g, yielding a net force of \frac{4}{3}\pi r^3 g (\rho_p - \rho). At low particle Reynolds numbers (Re_p < 1), where viscous forces dominate, the drag force follows Stokes' law: F_d = 6\pi \mu r v_t. Balancing this with the net gravitational force gives the terminal velocity v_t = \frac{2 r^2 g (\rho_p - \rho)}{9 \mu}. This expression, derived by George Gabriel Stokes in 1851, applies to laminar flow regimes typical of fine particles like silt in water. For high particle Reynolds numbers (Re_p > 10^3), inertial forces prevail, and the drag force is F_d = \frac{1}{2} C_d \rho v_t^2 \pi r^2, where C_d is the drag coefficient that becomes approximately constant (C_d \approx 0.44 for spheres). Balancing forces then yields v_t \approx \sqrt{\frac{4 g d (\rho_p - \rho)}{3 \rho C_d}}, with particle diameter d = 2r. The particle Reynolds number is defined as Re_p = \frac{\rho v_t d}{\mu} and is used iteratively to select the appropriate regime and C_d(Re_p) value, as C_d decreases with increasing Re_p in transitional regimes ($1 < Re_p < 10^3). In engineering applications, such as sedimentation tanks for water treatment, Re_p determines the settling regime to optimize particle removal efficiency; for instance, laminar conditions (Re_p < 1) are targeted for fine floc particles using to size tank overflow rates below v_t. In natural environments, raindrops of 1–3 mm diameter achieve terminal velocities of 6–9 m/s at high Re_p (200–3000), where the constant-C_d approximation governs their fall through air.

Engineering Applications

Pipe friction

In pipe flows, the Reynolds number plays a central role in determining frictional pressure losses, which are essential for engineering design in fluid transport systems. The Darcy-Weisbach equation quantifies the pressure drop ΔP due to wall friction over a pipe length L as \Delta P = f \frac{L}{D} \frac{\rho v^2}{2}, where D is the pipe diameter, ρ is the fluid density, v is the mean velocity, and f is the dimensionless Darcy friction factor. This factor f depends primarily on the Re = ρ v D / μ (with μ the dynamic viscosity) and the relative roughness ε/D, where ε represents the average height of surface protrusions on the pipe wall. For laminar flow at low Reynolds numbers (typically Re < 2300), the friction factor follows the analytical relation derived from the : f = \frac{64}{\mathrm{Re}}. This explicit form arises from the parabolic velocity profile in fully developed laminar pipe flow, where viscous forces dominate and pressure loss scales inversely with Re. In contrast, turbulent flow at higher Reynolds numbers (Re > 4000) requires empirical correlations for f, as inertial effects and wall roughness introduce complex interactions. The widely used provides an implicit solution: \frac{1}{\sqrt{f}} = -2 \log_{10} \left( \frac{\varepsilon / D}{3.7} + \frac{2.51}{\mathrm{Re} \sqrt{f}} \right), originally developed from experimental data on roughened pipes. This equation captures the transition from smooth-wall behavior (where f decreases with increasing Re) to fully rough regimes (where f becomes independent of Re and depends only on ε/D). The interplay between Re and ε/D is graphically represented in the Moody diagram, a log-log plot of f versus Re for various relative roughness values, enabling engineers to select f without iterative calculations for many practical cases. At low Re, curves for different ε/D converge due to laminar dominance; at high Re in the turbulent regime, they asymptote to horizontal lines for fully rough flow, highlighting how roughness amplifies friction beyond a critical Re threshold. For smooth pipes (ε ≈ 0), f continues to decline logarithmically with Re even at very high values, as confirmed by high-pressure experiments up to Re ≈ 10^6. In turbulent pipe flows, the average energy dissipation rate per unit mass ε, which quantifies irreversible losses due to , scales as ε = f v^3 / (2 D). This relation links directly to the friction factor and flow kinematics; for smooth s at high Re, where f ∝ [log_{10} (Re / 7)]^{-2}, ε exhibits weak dependence on Re beyond the primary v^3 / D scaling, with concentrated near the wall in a thin viscous sublayer whose thickness decreases as Re^{-1}. Experimental measurements in large-scale facilities confirm this , showing that bulk approaches an inviscid at extreme Re > 10^5, while total losses remain governed by wall shear.

Packed bed

In packed beds, which consist of a collection of solid particles forming a , the Reynolds number characterizes the flow regime and for fluids passing through the voids. These systems are prevalent in for processes requiring high surface area contact between fluid and solids, such as and particle separation. The particle Reynolds number, Re_p, adapted for packed beds, is defined as Re_p = \frac{\rho v d_p}{\mu (1 - \varepsilon)}, where \rho is the , v is the superficial velocity (volumetric flow rate divided by the total cross-sectional area), d_p is the equivalent particle , \mu is the dynamic , and \varepsilon is the bed porosity (void ). This formulation incorporates the solid (1 - \varepsilon) to reflect the effective flow obstruction by particles. Pressure drop across a packed bed is commonly predicted using the , an empirical correlation that separates viscous and inertial contributions: \frac{\Delta P}{L} = 150 \frac{\mu (1 - \varepsilon)^2 v}{\varepsilon^3 d_p^2} + 1.75 \frac{\rho (1 - \varepsilon) v^2}{\varepsilon^3 d_p}. The first term dominates at low Re_p, representing laminar viscous losses akin to , while the second term accounts for inertial effects at higher Re_p, arising from form drag on particles. This , derived from experiments on air and water flow through sands and beads, remains the standard for design despite minor variations in coefficients for non-spherical particles. Flow behavior in packed beds transitions through distinct regimes based on Re_p. In the Darcy regime (Re_p < 1), flow is purely viscous and linear with velocity, allowing simple permeability-based predictions without inertial corrections. The Forchheimer regime follows, where inertial effects augment the pressure drop nonlinearly, often modeled by adding a quadratic term to Darcy's law; this occurs for $1 < Re_p < 1000. At Re_p > 1000, turbulence emerges, with enhanced mixing and higher energy dissipation, though the bed structure suppresses full disorder compared to open-channel flows. These transitions influence scalability in designs, as inertial and turbulent effects increase pumping requirements. Packed beds find extensive use in fixed-bed chemical reactors, where controlled Re_p ensures uniform reactant distribution over catalysts for reactions like ammonia synthesis, and in filtration units, such as granular media filters for , where low Re_p minimizes particle displacement while capturing contaminants. In both, the Reynolds number guides optimization of bed depth, particle size, and flow rates to balance efficiency and energy use.

Stirred vessel

In stirred vessels, the mixing Reynolds number characterizes the flow regime during agitation processes and is defined as Re_m = \frac{\rho N D^2}{\mu}, where \rho is the fluid density, N is the impeller rotational speed, D is the impeller diameter, and \mu is the dynamic . This dimensionless parameter represents the ratio of inertial to viscous forces within the , enabling scale-up predictions for mixing operations in applications. Flow in stirred tanks transitions through distinct regimes based on Re_m: dominates for Re_m < 10, where viscous effects control the motion and result in smooth, layered streamlines; a occurs between $10 \leq Re_m \leq 10^4, featuring intermittent instabilities; and turbulent flow prevails for Re_m > 10^4, with inertial forces driving chaotic eddies and enhanced bulk circulation. These regimes influence performance, with geometry-specific variations in the exact transition points. The power number, a dimensionless measure of energy input defined as P_o = \frac{P}{\rho N^3 D^5}, where P is the power consumption, remains nearly in the turbulent regime (typically P_o \approx 5 for a standard Rushton turbine), indicating independence from . In contrast, during , P_o varies inversely with Re_m (P_o \propto 1/Re_m), highlighting the dominance of viscous dissipation. Blend time, the duration required to achieve compositional uniformity in the , and circulation patterns are strongly dependent on Re_m. In the turbulent regime, blend time is independent of and scales inversely with impeller speed (\theta_m \propto 1/N), promoting efficient radial and axial flows for rapid homogenization. In laminar conditions, higher prolongs blend time and restricts circulation to slower, viscosity-governed paths near the , while the transitional regime shows gradual improvements in mixing efficiency with increasing Re_m.

Advanced Concepts

Similarity of flows

The principle of in relies on achieving geometric, kinematic, and dynamic similarity between a model and its to ensure that behaviors predictably. Geometric similarity requires proportional scaling of all lengths, while kinematic similarity demands matching fields and streamlines. Dynamic similarity, which governs force ratios such as inertial to viscous effects, necessitates equal dimensionless parameters like the Reynolds number (Re) between the model and to replicate viscous influences accurately. In model testing, the Reynolds number is crucial for scaling viscous-dominated flows, often requiring adjustments to speed, fluid properties, or model size to match the prototype's Re. For instance, ship hull models in towing tanks are typically tested to satisfy the Froude number for wave patterns but at lower Re due to scale reduction, necessitating empirical corrections for viscous drag differences. To achieve closer Re matching, larger models or alternative fluids (e.g., higher-viscosity liquids) may be used, though practical constraints often limit full similitude. A key limitation arises in free-surface flows, where viscous effects (governed by ) conflict with gravitational effects (governed by the , ), making simultaneous matching impossible without specialized fluids or approximations. In ship model tests, for example, maintaining with water ensures wave similarity, but the resulting low underpredicts turbulent boundary layers, requiring post-test adjustments like ITTC correlation lines to estimate full-scale performance. This trade-off highlights the challenge in achieving complete dynamic similarity for gravity-influenced systems. In computational fluid dynamics (CFD), the Reynolds number dictates mesh resolution requirements for accurate turbulence modeling, as higher Re demands finer grids near walls to resolve thin boundary layers and capture inertial-viscous interactions without excessive numerical diffusion. For turbulent flows, Re influences the choice between Reynolds-averaged Navier-Stokes (RANS) models, which tolerate coarser meshes for high-Re simulations, and large eddy simulations (LES), which require grid sizes scaling inversely with Re to resolve energy-containing eddies. The Reynolds number thus guides discretization strategies to balance computational cost and fidelity in scaling virtual prototypes.

Smallest scales of turbulent motion

In fully developed turbulence at high Reynolds numbers, the smallest scales of motion are characterized by the Kolmogorov microscale, denoted as \eta, which represents the length scale at which viscous dissipation dominates and is converted into . This scale is given by \eta = \left( \frac{\nu^3}{\epsilon} \right)^{1/4}, where \nu is the kinematic viscosity and \epsilon is the mean rate of energy dissipation per unit mass. The Kolmogorov microscale arises from the assumption of local and homogeneity at small scales, independent of the larger flow structures, provided the Reynolds number is sufficiently large to allow a wide separation of scales. A key parameter characterizing the intensity of turbulence at these intermediate scales is the Taylor Reynolds number, Re_\lambda = \frac{u' \lambda}{\nu}, where u' is the root-mean-square fluctuation and \lambda is the , defined as the scale over which velocity correlations decay. This number, often on the order of hundreds or more in high-Re flows, quantifies the ratio of inertial to viscous effects near the energy-containing eddies, helping to delineate the regime where begins to influence the flow dynamics. Between the large -containing eddies and the dissipative Kolmogorov scales lies the inertial subrange, where cascades from larger to smaller eddies through nonlinear interactions without from , provided the Reynolds number greatly exceeds unity. In this subrange, the spectrum E(k) follows Kolmogorov's famous -5/3 , E(k) \propto \epsilon^{2/3} k^{-5/3}, with k being the , reflecting a universal statistical equilibrium independent of \nu. This cascade process ensures that the input at large scales is transferred conservatively down to the dissipative scales. The global Reynolds number plays a crucial role in determining the extent of scale separation in turbulent flows; as Re increases, the ratio of the integral length scale L (associated with large eddies) to \eta scales as Re^{3/4}, widening the inertial subrange and allowing more universal behavior at small scales. For instance, in turbulent pipe flows at high Re, this separation can span orders of magnitude, enabling the observation of the inertial subrange in experiments and simulations. Higher Re thus enhances the range over which viscosity-independent physics governs the turbulence microstructure. Direct numerical simulations (DNS) of must resolve down to the Kolmogorov to capture the full dynamics accurately, requiring a computational that scales with the number of points N \sim Re^{9/4} in three dimensions, due to the \eta \sim Re^{-3/4} dependence and the need to cover the domain volume. This steep scaling poses significant computational challenges for high-Re flows, limiting DNS to moderate Re in practice and motivating subgrid-scale modeling in large-eddy simulations.

In physiology

In physiological systems, the plays a crucial role in characterizing of biological flows, particularly in the cardiovascular and respiratory systems, where it helps predict whether flows remain laminar or transition to turbulent regimes. In circulation, the Reynolds number typically ranges from approximately 1 to 4000 in arteries, reflecting the pulsatile nature of cardiac-driven flow that generally maintains laminar conditions despite periodic peaks during . This range ensures efficient transport of oxygen and nutrients without excessive energy dissipation, as higher values up to approximately 4000 in larger arteries like the can introduce mild disturbances but rarely full under normal conditions. In the microcirculation, such as capillaries, the Reynolds number drops below 1—often to 0.001–0.01—due to the small vessel diameters (around 5–10 μm) and low velocities, promoting purely laminar flow that facilitates passive diffusion of gases and solutes across vessel walls without disruptive mixing. This low Reynolds regime in capillaries and venules minimizes shear stresses on endothelial cells, thereby preventing potential damage from turbulent eddies or high-velocity fluctuations that could otherwise compromise vascular integrity. At the , peak Reynolds numbers exceeding 2000 during ejection promote enhanced mixing of blood in the , aiding in the uniform distribution of nutrients and hormones through disturbed but controlled flow patterns rather than outright . In the , airflow in the larger bronchi exhibits Reynolds numbers of approximately 1000–4000, where transitional flow occurs due to branching and higher velocities, but it shifts to fully laminar conditions (Re < 1000) in smaller airways and alveoli, optimizing gas exchange by maintaining stable boundary layers. Evolutionary and physiological adaptations in vascular architecture, such as the gradual tapering of arteries, help sustain optimal Reynolds numbers along the vascular tree by balancing diameter reduction with flow velocity increases, thereby stabilizing laminar flow and reducing the risk of turbulence in distal segments. This tapering design enhances transport efficiency while adapting to varying hemodynamic demands across organ systems.

Complex systems

In complex systems beyond traditional fluids, the Reynolds number serves as an analogy for dimensionless parameters that quantify the balance between inertial-like (propagative or momentum-driven) forces and dissipative or interactive forces, often marking transitions from ordered to disordered states akin to laminar-to-turbulent shifts in fluids. In traffic flow models, an analogous "Reynolds number" is formulated as the ratio of inertial forces—arising from vehicle density and average speed—to "viscous" forces representing inter-vehicle interactions and road friction. This parameter, termed the Traffic Flow Factor (TFF), highlights jamming transitions at critical densities where flow capacity peaks before congestion emerges, mirroring turbulence onset. Such transitions occur when vehicle density exceeds approximately 20-30 vehicles per kilometer per lane on highways, leading to sharp drops in flow rate. Similar Re-like parameters appear in neural networks, particularly in analyzing signal propagation through recurrent architectures. Here, the product of network connectivity strength and synaptic gain acts as a control parameter for dynamical regimes: low values yield stable, ordered propagation suitable for reliable computation, while values above a critical threshold (typically around 1 for balanced excitation-inhibition) induce chaotic activity, enhancing computational expressivity but risking signal divergence. This bifurcation parallels high-Reynolds-number instability in fluids, with chaos enabling richer pattern recognition in tasks like . In granular flows, the granular Reynolds number Re_{gr} = \frac{\rho d^{2} \dot{\gamma}}{\mu_{eff}}, where \rho is particle density, d is particle diameter, \dot{\gamma} is the shear rate, and \mu_{eff} is effective viscosity, embodies by delineating flow regimes. At low Re_{gr} (quasi-static regime, Re_{gr} \ll 1), frictional contacts dominate, yielding slow, nearly rigid-like motion as in silos or avalanches. High Re_{gr} (inertial regime, Re_{gr} \gg 1) shifts to collision-driven dynamics, with stress scaling quadratically with shear rate (\tau \propto \dot{\gamma}^2), as observed in rapid dense flows like hourglass discharge. This separation, rooted in Bagnold's 1954 experiments, predicts rheological behavior across scales from lab tests to geophysical events. While these analogies illuminate transition dynamics, the classical Reynolds number does not strictly apply outside Newtonian fluids, as underlying mechanisms differ (e.g., discrete collisions in granulars versus continuous viscosity). Nonetheless, analogous dimensionless groups reliably capture similar bifurcations, fostering cross-disciplinary insights into stability and disorder.

Relationship to other dimensionless parameters

The Reynolds number () interacts with the (M) in compressible fluid flows, where M characterizes the ratio of flow speed to the speed of sound, indicating effects that become prominent when M > 0.3. In contrast, governs the relative importance of inertial to viscous forces independently of , allowing separate analysis of viscous dissipation in high-speed regimes. For instance, in turbulent layers, normalized Reynolds stresses remain largely independent of even at elevated Mach numbers, underscoring Re's distinct role in viscous scaling. influences structure primarily when the root-mean-square Mach number approaches or exceeds 1, altering energy transfer mechanisms while continues to dictate the onset of . In free-surface flows, such as those involving ships or open channels, Re conflicts with the (Fr = v / \sqrt{g L}), which measures the ratio of inertial to gravitational forces and is essential for wave pattern similitude. Dynamic similarity requires matching both Re and Fr between model and , but scaling down length reduces Re disproportionately, leading to viscous effects that do not replicate full-scale behavior. This incompatibility necessitates distorted models, where geometric scales differ vertically and horizontally to approximate both parameters, or empirical corrections like form factors to adjust resistance predictions. In ship testing, for example, Froude-based scaling prioritizes wave resistance, with Reynolds effects mitigated through stimulation or computational adjustments. In convective , Re combines with the (Pr = \nu / \alpha), the ratio of momentum to , in (Nu) correlations for , typically expressed as Nu = f(Re, Pr), where higher Re promotes turbulent mixing and enhances heat transfer rates. For internal flows like pipes, empirical relations such as Nu \approx 0.023 Re^{0.8} Pr^{0.4} apply under turbulent conditions (Re > 10^4), emphasizing Re's control over flow regime. In mixed convection scenarios, the (Gr), representing the ratio of buoyancy to viscous forces, integrates with Re and Pr; for instance, Nu increases with Re across Prandtl values but shows weaker dependence on Gr when dominates, as in vertical channels with aiding flows. The (Pe = Re \cdot Pr) extends Re's influence by quantifying the competition between advective transport and molecular diffusion in heat or mass transfer processes. When Pe > 1, advection prevails, streamlining scalar fields and reducing diffusive spreading, as seen in high-Re flows with moderate Pr. This parameter is particularly relevant in low-Prandtl fluids, like liquid metals, where Pe highlights regimes dominated by convective over conduction.

References

  1. [1]
    The Reynolds Number: A Journey from Its Origin to Modern ... - MDPI
    Some of the basic applications of the Reynolds number are the study of flow dynamics in pipes, aircraft design, automotive, heat exchangers, industrial mixers, ...
  2. [2]
    This Month in Physics History | American Physical Society
    March 15, 1883: Osborne Reynolds Proposes the Reynolds Number. The intricate motion of fluids is notoriously difficult to predict, due to these substances' ...
  3. [3]
    III. An experimental investigation of the circumstances which ...
    Cite this article. Reynolds Osborne. 1883III. An experimental investigation of the circumstances which determine whether the motion of water shall be direct ...
  4. [4]
    Reynolds Number
    The Reynolds number expresses the ratio of inertial (resistant to change or motion) forces to viscous (heavy and gluey) forces.
  5. [5]
    Dimensional Analysis and Similarity
    The Buckingham Pi technique is a formal "cookbook" recipe for determining the dimensionless parameters formed by a list of variables. There are six steps, which ...
  6. [6]
    Dimensional Analysis – Introduction to Aerospace Flight Vehicles
    The Reynolds number is one of the most significant parameters in aerodynamics because it governs the relative magnitude of viscous effects to inertia effects in ...<|separator|>
  7. [7]
    [PDF] 5.5 Buckingham Pi theorem - MIT
    May 13, 2010 · The Buckingham Pi theorem provides that number. I derive it with a series of examples. Here is a possible beginning of the theorem statement: ...
  8. [8]
    [PDF] Nondimensionalization of the Navier-Stokes Equation
    Nondimensionalization of the Navier-Stokes equation involves using scaling parameters to define nondimensional variables and multiplying by L/(ρV) to cancel ...
  9. [9]
    [PDF] Lecture 12: Scaling and Nondemensionalization
    Scaling and nondimensionalization are tools to analyze equations by scaling variables and parameters to make terms order one, creating nondimensional equations.
  10. [10]
    [PDF] Reynold's transport theorem
    This is the non-dimensional form of the Navier-Stokes equation and the non-dimensional group,. ρUL/η is known as the Reynolds number, Re. Du∗. Dt∗. = −∇∗P ...
  11. [11]
    [PDF] 1 Introduction. 2 Boundary Layer Governing Equations. - MIT
    If we define the Reynolds number as R = umaxδ/nu, equation. (8.177) implies that R ∝ x1/3. To illustrate the streamlines for the flow generated by the jet, we ...
  12. [12]
    [PDF] Boundary layers
    The. Reynolds number is as usual calculated as Re ≈ UL/ν where L is a typical length scale for significant changes in the flow, determined by the geometry of.
  13. [13]
    Mach Number & Reynolds Number – Introduction to Aerospace ...
    Definition of Reynolds Number​​ Notice that when quoting the value of the Reynolds number, the characteristic length must be defined, e.g., the Reynolds number ...
  14. [14]
    [PDF] Fluid flow in pipes of rectangular cross sections - Scholars' Mine
    hydraulic radius is the proper length dimension in Reynolds number subject ... as the pipe diameter, D, defining Reynolds number as VDP. Character ...
  15. [15]
    [PDF] 7. Basics of Turbulent Flow - MIT
    Given the characteristic velocity scale, U, and length scale, L, for a system, the Reynolds number is Re = UL/ν, where ν is the kinematic viscosity of the fluid ...
  16. [16]
    [PDF] The History of Poiseuille's Law
    In 1839, a German hydraulic engineer, Gotthilf Heinrich Ludwig Hagen ... laminar flow law should be called the Hagen-Poiseuille law as advocated by ...
  17. [17]
    XXIX. An experimental investigation of the circumstances which ...
    An experimental investigation of the circumstances which determine whether the motion of water shall be direct or sinuous, and of the law of resistance in ...
  18. [18]
    [PDF] Note on the History of the Reynolds Number
    Actually Sommerfeld's work is not foremost in one's mind when one thinks of the direct continuation of the ideas set forth by Osborne Reynolds in 1883 (Reynolds ...
  19. [19]
    Transition and Turbulence - Princeton University
    While the transition from laminar to turbulent flow occurs at a Reynolds number of approximately 2300 in a pipe, the precise value depends on whether any small ...
  20. [20]
    Critical Reynolds Number for a Natural Transition to Turbulence in ...
    Feb 6, 2007 · Thus, the Reynolds number of ≈ 2000 has been referred to in the literature as the “critical” Reynolds number ( Re cr ) for a “natural“ ...Abstract · Article Text
  21. [21]
    Laminar, Transitional and Turbulent Flow - The Engineering ToolBox
    The Reynolds number is proportional to inertial force divided by viscous force. The flow is. laminar when Re < 2300; transient when 2300 < Re < 4000; turbulent ...
  22. [22]
    Tollmien-Schlichting Wave - an overview | ScienceDirect Topics
    Tollmien-Schlichting waves are defined as primary instabilities that lead to the transition from laminar to turbulent flow in two-dimensional and ...
  23. [23]
    [PDF] bypass transition to turbulence and research desiderata
    Bypass transitions are seldom mentioned in texts or meetings on instabil- ity and transition to wall turbulence. Like poor relations, they are untidy;.
  24. [24]
    [PDF] A review of factors affecting boundary-layer transition
    Some of these factors are longitudinal pressure gradients, lateral pressure gradients resulting from three-dimensional-flow effects, the ratio of surface ...
  25. [25]
    [PDF] Modeling Of Pipe Flows And Observation Of Laminar Turbulent ...
    The Reynolds number of transition was demonstrated to vary from the accepted value of 2300, depending on tube inlet geometry.
  26. [26]
    Poiseuille Flow - Thermopedia
    The volumetric flow rate through the pipe is given by. (4). This is the Hagen-Poiseuille Equation, also known as Poiseuille's Law. Experimentally, Eq. (4) is ...
  27. [27]
    [PDF] Turbulent Pipe Flows
    This implies blunter velocity profiles as the Reynolds number is increased and this is exemplified by the sample velocity profiles shown in Figure 3. When the ...
  28. [28]
    "Friction Factors for Pipe Flow" (Moody, Lewis F., 1944, Trans. ASME ...
    The paper was intended for application to normal conditions of engineering practice and specifies a number of qualifications limiting the scope of the charts, ...
  29. [29]
    Fluid Flow - Entrance Length and Developed Flow
    The entrance length is the length in a tube or duct after an obstruction - until the flow velocity profile is fully developed. A fluid flow need some length to ...
  30. [30]
    Equivalent Reynolds Number - an overview | ScienceDirect Topics
    For rectangular flow channels, the hydraulic diameter is defined. (3) d h = 4 ab 2 ( a + b ). There are several different definitions of the channel aspect ...<|control11|><|separator|>
  31. [31]
    Aspect Ratio Effects on Turbulent and Transitional Flow in ...
    As the Reynolds number is further increased, the measured velocity near the channel centerline deviates further from the laminar prediction and becomes smaller, ...
  32. [32]
    [PDF] Turbulent Friction in Rectangular Ducts
    It was found that at constant Reynolds number based on hydraulic diameter the friction factor increases monotonically with increasing aspect ratio.
  33. [33]
    Effect of aspect ratio on the laminar-to-turbulent transition in ...
    The results show that the critical Reynolds number decreases with the increasing aspect ratio. However, the relative location of laminar breakdown does not ...
  34. [34]
    Effect of aspect ratio and inlet manifold shape on the laminar-to ...
    Mar 1, 2021 · Results showed that sudden contraction shows the most delayed transition at around Reynolds number (Re) of 2000 whereas swirl inlet type showed ...
  35. [35]
    (PDF) Specific aspects of turbulent flow in rectangular ducts
    Secondary flows of the first kind are also called skew-induced or pressure-driven secondary flows. These flows arise in both laminar and turbulent flow regimes ...
  36. [36]
    [PDF] Signature redacted - DSpace@MIT
    An analysis is presented by which it is shown that secondary flows cannot exist in fully developed laminar flow through a straight tube of any cross-sectional ...
  37. [37]
    Secondary flow in turbulent ducts with increasing aspect ratio
    May 17, 2018 · Direct numerical simulations of turbulent duct flows with aspect ratios 1, 3, 5, 7, 10, and 14.4 at a center-plane friction.<|control11|><|separator|>
  38. [38]
    On the role of secondary motions in turbulent square duct flow
    May 24, 2018 · Secondary motions are found to contribute approximately 6 % of the total friction, and to act as a self-regulating mechanism of turbulence ...<|control11|><|separator|>
  39. [39]
    Fast and accurate prediction of airflow and drag force for duct ...
    Aug 15, 2018 · Rectangular ducts are commonly adopted in ventilation systems to push airflow through pipes, with Reynolds number reaching up to 500000.
  40. [40]
    [PDF] BASIC HYDRAULIC PRINCIPLES OF OPEN-CHANNEL FLOW
    where the hydraulic radius, R, is defined as the area divided by the wetted ... which means the Froude number, Fr, is equal to 1 at critical flow. Finally ...
  41. [41]
    On the classification of open channel flow regimes, Dr. Victor M. Ponce
    The Reynolds number R = ud /ν (u = mean velocity, d = flow depth, and ν = kinematic viscosity) is used to characterize laminar or turbulent flow, a small R ...
  42. [42]
    [PDF] Laws of turbulent flow in open channels
    pipes for a given Reynolds number R. is not affected by the shape of the pipe, provided that in forming the Reynolds number the hydraulic radius R is used ...
  43. [43]
    Basic Understanding of Airfoil Characteristics at Low Reynolds ...
    As the Reynolds number decreases, there is an increase in drag (again, possibly because of premature separation of the laminar boundary layer).
  44. [44]
    [PDF] Low Reynolds Number Airfoil Design Lecture Notes
    Only single-element airfoils are considered here. Airfoils for such aircraft typically operate in the Reynolds number range 200,000 to. 500,000. For example ...
  45. [45]
    [PDF] .IAfAA - UIUC Applied Aerodynamics Group
    At low Reynolds numbers, the presence of laminar separation bubbles often results in the increased drag seen on some airfoils. In an effort to mitigate these ad ...
  46. [46]
    Investigation on flow structure and aerodynamic characteristics over ...
    May 14, 2021 · This Review summarizes the progress in research on the flow structure and aerodynamic characteristics of an airfoil at a low Reynolds number<|control11|><|separator|>
  47. [47]
    Boundary Layer
    For lower Reynolds numbers, the boundary layer is laminar and the streamwise velocity changes uniformly as one moves away from the wall, as shown on the left ...
  48. [48]
    Aerodynamics of Airfoil Sections – Introduction to Aerospace Flight ...
    However, as the Reynolds number decreases below a million, the lift and drag curves become significantly more rounded and eventually become nonlinear for ...
  49. [49]
    Investigation of Reynolds Number Effects on Aerodynamic ... - MDPI
    Jul 1, 2021 · The results show that the Reynolds number has a significant impact on the boundary layer displacement thickness, surface pressure distribution, shock wave ...
  50. [50]
    Scale Effect on Clark Y Airfoil Characteristics from NACA Full-Scale ...
    With increasing Reynolds number the airfoil characteristics are affected in the following manner: the drag at zero lift decreases, the maximum lift increases, ...
  51. [51]
    [PDF] Wind Tunnel Testing Airfoils at Low Reynolds Numbers
    This paper describes the wind tunnel testing methodology that has been applied to testing over 200 airfoils at low Reynolds numbers (40,000 to 500,000).
  52. [52]
    [PDF] Scaling Techniques Using CFD and Wind Tunnel Measurements for ...
    This thesis addresses scaling aerodynamic data from wind tunnels to free flight, using CFD to correct for model support, wall interference, and lower Reynolds ...<|control11|><|separator|>
  53. [53]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · At this Reynolds number the boundary layer transitions from laminar to turbulent on the sphere surface and, thus, separates further downstream.<|separator|>
  54. [54]
    [PDF] Chapter9 ExternalFlow Chapter 9 External Flow Past Bodies
    The drag coefficient decreases when the drag coefficient decreases when the boundary layer becomes turbulent boundary layer becomes turbulent. Page 99. Flow ...
  55. [55]
    Drag of Blunt Bodies and Streamlined Bodies
    The boundary layer over the front face of a sphere or cylinder is laminar at lower Reynolds numbers, and turbulent at higher Reynolds numbers. When it is ...
  56. [56]
    None
    Nothing is retrieved...<|control11|><|separator|>
  57. [57]
    Bluff Body Flows – Introduction to Aerospace Flight Vehicles
    Those with sharp, angular shapes tend to have fixed points of flow separation and drag coefficients that are insensitive to the Reynolds number.
  58. [58]
    On the flow past a sphere at low Reynolds number | Journal of Fluid ...
    Mar 29, 2006 · The flow of an incompressible, viscous fluid past a sphere is considered for small values of the Reynolds number. In particular the drag is ...
  59. [59]
    [PDF] Oseen Flow
    0. When the drag on the sphere in Oseen flow is evaluated it transpires that. Drag = 6πμRU. 1 +. 3Re. 16. (Blh4). However, because the flow close to the surface ...
  60. [60]
    Terminal Velocity: Stokes' Legacy in Fluvial Hydraulics
    Aug 28, 2019 · In this paper, Stokes made the first breakthrough in calculating the fluid drag—also called Stokes' law, which defines the Stokes drag FD on ...
  61. [61]
    Terminal fall velocity: the legacy of Stokes from the perspective of ...
    In this paper, Stokes made the first breakthrough in calculating the fluid drag ... Stokes drag FD on a spherical particle of diameter d as. F D = 3 π μ U d ...Missing: original | Show results with:original
  62. [62]
    Drag of a Sphere | Glenn Research Center - NASA
    Jun 30, 2025 · Drag of a Sphere. On this page: Drag Coefficient; Cases of Flow Past a Cylinder and a Sphere; Experimental Observations of Reynolds Number.Missing: empirical | Show results with:empirical
  63. [63]
    Dependence of square cylinder wake on Reynolds number
    Jan 5, 2018 · In the laminar regime (i.e., at Re < Rec1), time-mean drag coefficient C ¯ d exponentially declines with Re.13,15,19 In the turbulent regime (Re > ...<|control11|><|separator|>
  64. [64]
    Strouhal numbers of rectangular cylinders | Journal of Fluid Mechanics
    Apr 20, 2006 · The results show how Strouhal number varies with a width-to-height ratio of the cylinders in the range of Reynolds number between 70 and 2 × l04 ...
  65. [65]
    Rectangular Cylinder Orientation and Aspect Ratio Impact on ... - MDPI
    It is found that the flow characteristics are highly influenced by changes in the aspect ratio compared to the Reynolds number. The flow exhibits three ...
  66. [66]
    Reynolds-number-effects in flow around a rectangular cylinder with ...
    The model was a sharp-edged rectangular cylinder with aspect ratio height/width 1:5 (width/span ratio 1:10.8), which was investigated in both basic orientations ...
  67. [67]
    [PDF] Summary of Drag Coefficients of Various Shaped Cylinders - DTIC
    Drag coefficients of the various shapes depend on both Reynolds number and Mach num- ber. At low air velocities the Reynolds number almost entirely determines ...
  68. [68]
    On the Reynolds number sensitivity of the aerodynamics of bluff ...
    A review of recent comparisons between the aerodynamic characteristics of bluff bodies with sharp edges at low and high Reynolds number is presented.
  69. [69]
    [PDF] Reynolds-Number-Effects in Flow around a rectangular Cylinder ...
    The results show that, there is indeed no systematic variation depending on the Reynolds number. The mean drag coefficient with Cd = 2.4 is roughly 10% higher ...
  70. [70]
    [PDF] on the Motion of Pendulums. By G. G. Stokes, M.A., Fellow of ...
    its terminal velocity depends almost wholly on the internal friction of air. ... The fifth section of the first part of the present paper contains an.
  71. [71]
    The Physics of Falling Raindrops in Diverse Planetary Atmospheres
    Mar 15, 2021 · Terminal velocity vT occurs when the raindrop is no longer accelerating, and the gravitational force Fg is balanced by the aerodynamic drag ...
  72. [72]
    [PDF] Sedimentation - TU Delft OpenCourseWare
    When the Reynolds number Re > 1,600, settling is turbulent and when 1<Re<1,600, settling is in transition between laminar and turbulent. In Figure 4 the ...
  73. [73]
    [PDF] Friction Factors for Pipe Flow
    Pigott (4) published a chart for the same friction factor, using the same co-ordinates as in Fig. 1 of this paper. His chart has proved to be most useful and ...
  74. [74]
    [PDF] (E) High Reynolds-Number Scaling of Channel & Pipe Flow
    It is therefore suggested that, at extremely high. Reynolds numbers, the “inviscid” energy dissipation in the bulk may become greater than the dissipation in ...
  75. [75]
    [PDF] Turbulent Pipe Flow at Extreme Reynolds Numbers
    Feb 28, 2012 · Here, we report turbulence measurements over an unprecedented range of Reynolds numbers using a unique combination of a high-pressure air ...
  76. [76]
    Fluid dynamics in a packed bed reactor - ScienceDirect.com
    Feb 1, 2022 · This research frames a novel approach to develop a pseudo-continuous model for describing the dynamics of a Newtonian fluid in an industrial-scale packed-bed ...
  77. [77]
    [PDF] Flow of Fluids through Granular Beds and Packed Columns
    Fluid flow through granular beds is common in chemical industry. Darcy's law describes flow as proportional to pressure and inversely to bed thickness.
  78. [78]
    Fluid flow through packed columns - Semantic Scholar
    Fluid flow through packed columns · S. Ergun · Published 1952 · Engineering, Chemistry, Materials Science.
  79. [79]
    From Darcy to turbulent flow: Investigating flow characteristics and ...
    Oct 29, 2024 · Flow regime characteristics in porous media flows at high Reynolds numbers ... Flow regimes in packed beds of spheres from predarcy to turbulent.
  80. [80]
    Impeller Reynolds Number - an overview | ScienceDirect Topics
    Impeller Reynolds number is defined as a dimensionless value that characterizes the flow regime in stirred vessels, calculated using the formula \( Re_i = N_i ...
  81. [81]
  82. [82]
  83. [83]
    [PDF] ChE 512
    Laminar and turbulent zones co-exist. Discharge flow increases and reaches maximum at transition Reynolds number (Re ≈ 90). Velocity of the liquid away from the ...
  84. [84]
    Mixing time in stirred vessels: A review of experimental techniques
    Mixing time can be expressed in its nondimensional form: N θ m = K where θm is the mixing time in s, N is the impeller speed in revolutions per second and K is ...
  85. [85]
    [PDF] Chapter 6 Mixing - ResearchGate
    stirred vessels make mixing performance poor. For mixing to be effective, fluid circulated by the impeller must sweep the entire vessel in a reasonable time.
  86. [86]
    [PDF] 26 Dec 2016 Chapter 07: Dimensional Analysis 5. Modeling and ...
    Dec 26, 2016 · Maintaining both Reynolds number and Froude number similarity is difficult to achieve in practice.Missing: principle | Show results with:principle
  87. [87]
    Dynamic Similarity – Introduction to Aerospace Flight Vehicles
    Navier-Stokes Equations. The scaling significance of the Reynolds number and the Mach number can also be determined from the momentum equation of fluid ...<|separator|>
  88. [88]
    [PDF] Chapter 7 Resistance and Powering of Ships
    This chapter will investigate the differing forms of hull resistance, ship power transmission, and the screw propeller. Additionally, we will investigate ship ...
  89. [89]
    [PDF] 7.5-04-05-01 Guideline on the determination of model-ship ... - ITTC
    The basis for the determination of these cor- relation factors is always the comparison be- tween predicted full scale values based on the towing tank ...
  90. [90]
  91. [91]
    A self-adapting turbulence model for flow simulation at any mesh ...
    Nov 12, 2007 · ... Reynolds number and also quickly damps the smallest scales of the turbulence. ... turbulence at any mesh resolution. The fundamental reason that ...<|separator|>
  92. [92]
  93. [93]
  94. [94]
    [PDF] The Local Structure of Turbulence in Incompressible Viscous Fluid ...
    Nauk SSSR (1941) 30(4). Paper received 28 December 1940. This translation by V. Levin, reprinted here with emendations by the editors of this volume. Proc ...
  95. [95]
    [PDF] Dissipation of Energy in the Locally Isotropic Turbulence Author(s)
    Paper received 30 April 1941. This translation by V. Levin, reprinted ... 28 of Kolmogorov (1941b) cannot be directly used for the determination of C ...
  96. [96]
    Influence of the spatial resolution on fine-scale features in DNS of ...
    ... DNS of homogeneous isotropic turbulence of dimension L3b can be estimated with the relation N ≈ (Lb/η)3 ∼ Re9/4Lb where R ⁢ e L b is the Reynolds ...
  97. [97]
    Vascular Arterial Haemodynamics - NCBI - NIH
    For an artery, the flow tends to change from laminar to turbulent at a Reynolds' number of approximately 2000.
  98. [98]
    [PDF] BLOOD FLOW IN ARTERIES - MRI Questions
    The typical Reynolds number range of blood flow in the body varies from 1 in small arterioles to approximately 4000 in the largest artery, the aorta.
  99. [99]
    Literature Survey for In-Vivo Reynolds and Womersley Numbers of ...
    Published Reynolds and Womersley numbers for 14 major arteries in the human body were determined via a literature search. Preference was given to in-vivo ...
  100. [100]
    [PDF] CHAPTER 5 - Blood Flow in Heart, Lung, Arteries, and Veins
    In the capillary blood vessels, the Reynolds number is in the order of 0.001 to 0.01, so small that it suggests complete insignificance of the inertial force.
  101. [101]
    Anatomy, Blood Flow - StatPearls - NCBI Bookshelf
    Reynold's number predicts the chances of flow being turbulent. The higher the number, the increased likelihood of being turbulent and vice versa. Reynold's ...
  102. [102]
    Hydrodynamics of Aortic Blood Flow
    As a substitute for a direct experimental approach, the Reynolds number has been used to predict the type of flow that may oc- cur in various parts of the aorta ...
  103. [103]
    Steady and oscillatory flow in the human bronchial tree
    For both cases, the peak Reynolds number and Womersley number are Re = 1500 and Wo = 12 , respectively.
  104. [104]
    Chaotic mixing deep in the lung - PNAS
    Because gas flow in the alveolar region occurs at very low Reynolds number (12) and the wall motion of the lung during breathing is essentially reversible (13– ...
  105. [105]
    Effect of vessel tapering on the transition to turbulent flow - PubMed
    The transition Reynolds number (based on the diameter of the tube at the point of measurement) increased as the angle of tapering increased.Missing: evolutionary | Show results with:evolutionary
  106. [106]
    Urban Traffic: Understanding the Traffic Flow Factor Through Fluid Dynamics
    ### Summary of Analogy Between Traffic Flow Factor (TFF) and Reynolds Number
  107. [107]
    [PDF] Traffic Flow Theory - A State-of-the-Art Report - ROSA P
    This is a state-of-the-art report on Traffic Flow Theory, covering traffic stream characteristics and human factors.
  108. [108]
    Transition to Chaos in Random Neuronal Networks | Phys. Rev. X
    Nov 19, 2015 · In this work, we investigate rate-based dynamics of neuronal circuits composed of several subpopulations with randomly diluted connections.
  109. [109]
    The Universal Presence of the Reynolds Number - MDPI
    He mentions that Arnold Sommerfeld introduced the term “Reynolds number” in hydrodynamics in 1908 at the Fourth International Congress of Mathematicians held ...