Fact-checked by Grok 2 weeks ago

Moody chart

The Moody chart, also known as the Moody diagram, is a logarithmic graph in that correlates the Darcy–Weisbach friction factor (f) with the (Re) for fully developed flow in circular , parameterized by curves representing different values of relative roughness (ε/D), where ε is the height of surface irregularities and D is the pipe diameter. This chart enables engineers to visually determine the required for calculating or head loss in pipe systems using the , h_f = f (L/D) (V² / 2g), where h_f is the head loss, L is the pipe length, V is the velocity, and g is . Developed by American hydraulic engineer Lewis F. Moody and first published in 1944, the chart synthesizes experimental data from numerous sources, including early 20th-century studies on turbulent by researchers such as Nikuradse and Prandtl, to provide a unified tool for practical applications in clean commercial pipes under steady, conditions. Moody's work built upon the implicit Colebrook–White equation, which relates f to and ε/D through the formula 1/√f = -2 log₁₀ [ (ε/(3.7D)) + (2.51/(Re √f)) ], by solving it iteratively to generate the chart's curves for turbulent regimes. The diagram distinguishes key flow regimes: (where f = 64/ for Re < 2,300), a transitional critical zone (2,300 < Re < 4,000), smooth-wall turbulent flow (dependent primarily on Re), and fully rough turbulent flow (where f is independent of Re and depends only on ε/D). Since its introduction, the Moody chart has become a cornerstone in chemical, mechanical, and civil engineering for designing piping networks, pumps, and hydraulic systems, offering a non-iterative approximation that avoids direct solution of the Colebrook equation while accounting for surface roughness effects from materials like drawn tubing (ε ≈ 0.0015 mm) to concrete (ε ≈ 0.3–3 mm). Its logarithmic scales—f on the vertical axis (typically 0.008 to 0.1) and Re on the horizontal (10² to 10⁸)—facilitate interpolation for intermediate values, though modern computational tools often supplement it for precision. Limitations include applicability to Newtonian fluids in steady flow and exclusion of entrance effects or non-circular ducts, for which modifications like hydraulic diameter adjustments are required.

Core Concepts in Pipe Flow

Reynolds Number

The Reynolds number (Re) is a dimensionless parameter fundamental to fluid mechanics, characterizing the flow regime in pipes and other conduits. It is defined as Re = \frac{\rho V D}{\mu} or equivalently Re = \frac{V D}{\nu}, where \rho is the fluid density, V is the average velocity of the fluid, D is the inner diameter of the pipe, \mu is the dynamic viscosity of the fluid, and \nu = \mu / \rho is the . This formulation arises from experimental observations in pipe flow, originally established by through systematic studies of water flow transitions in tubes of varying diameters. Physically, the Reynolds number quantifies the ratio of inertial forces to viscous forces acting on the fluid elements. Inertial forces scale with \rho V^2, promoting flow instability and mixing, while viscous forces scale with \mu V / D, which dampen disturbances and maintain orderly motion. When Re is low (below approximately 2300), viscous forces dominate, resulting in where fluid particles move in smooth, parallel layers. As Re increases to the range of 2300 to 4000, the flow enters a transitional regime with intermittent instabilities. Above 4000, inertial forces prevail, leading to characterized by chaotic eddies and enhanced mixing. These critical values are approximate and can vary slightly with entrance conditions or pipe disturbances, but they provide a reliable indicator for pipe flows. The Reynolds number originates from the non-dimensionalization of the Navier-Stokes equations, which govern viscous incompressible flow in pipes. The process begins with the dimensional momentum equation: \rho \left( \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} \right) = -\nabla p + \mu \nabla^2 \mathbf{u} + \rho \mathbf{g}. Non-dimensional variables are introduced using characteristic scales: position \mathbf{x}^* = \mathbf{x} / D, velocity \mathbf{u}^* = \mathbf{u} / V, pressure p^* = (p - p_\infty) / (\rho V^2), and time t^* = t / (D / V). Substituting these into the equation and dividing through by the inertial scale \rho V^2 / D yields the non-dimensional form: \frac{\partial \mathbf{u}^*}{\partial t^*} + (\mathbf{u}^* \cdot \nabla^*) \mathbf{u}^* = -\nabla^* p^* + \frac{1}{Re} \nabla^{*2} \mathbf{u}^* + \mathbf{g}^*, where Re = \rho V D / \mu appears as the coefficient balancing the nonlinear inertial terms against the viscous diffusion term. This scaling reveals Re as the key parameter dictating whether viscous effects can stabilize the flow against inertial instabilities in pipe geometries. As a dimensionless quantity, the Reynolds number carries no units, enabling direct comparisons across scales, fluids, and conditions. For typical engineering applications with water at 20°C (\nu \approx 10^{-6} m²/s), Re values in pipes often span $10^4 to $10^6. For instance, water flowing at 2 m/s in a 5 cm diameter pipe yields Re ≈ $10^5, firmly in the turbulent regime common to commercial piping systems. The Reynolds number is essential as a prerequisite for predicting the friction factor in pipe flow, as it identifies the governing regime for head loss correlations. In the context of the Moody chart, Re serves as the primary abscissa to delineate laminar, transitional, and turbulent regions.

Relative Roughness

Relative roughness, denoted as \epsilon / D, is a dimensionless parameter that characterizes the roughness of a pipe's internal surface relative to its diameter. It is defined as the ratio of the absolute roughness \epsilon, representing the average height of microscopic surface irregularities, to the internal diameter D of the pipe. The absolute roughness \epsilon originates from the pipe's material composition and manufacturing processes, such as drawing, casting, or galvanizing, which introduce inherent surface irregularities. Different pipe materials exhibit characteristic ranges of \epsilon values, influencing flow resistance in engineering applications. Typical values for common materials are summarized below:
MaterialAbsolute Roughness \epsilon (mm)
Drawn tubing0.0015
Commercial steel0.045
Galvanized iron0.15
Cast iron0.25
These values are based on empirical measurements and are widely used in design handbooks. Absolute roughness is measured using profilometry techniques, which employ a stylus to trace surface profiles and quantify height deviations, or non-contact optical and laser methods for precision assessment. Empirical values are also derived from standardized engineering data, such as those in , which provides guidelines for evaluating surface texture including roughness parameters. In turbulent pipe flow, relative roughness governs the interaction between surface protrusions and the near-wall boundary layer. Roughness elements that extend beyond the viscous sublayer disrupt laminar flow structures, promoting enhanced mixing and shear stress at the wall, thereby increasing frictional drag. In the fully rough regime, occurring at sufficiently high Reynolds numbers, the influence of viscosity diminishes, rendering the friction factor independent of the Reynolds number and dependent only on \epsilon / D. This regime was experimentally demonstrated in Johann Nikuradse's 1933 studies on artificially roughened pipes using sand grains, which established the foundational understanding of roughness effects in turbulence.

Darcy Friction Factor

The Darcy friction factor, denoted as f, is a dimensionless coefficient that quantifies the frictional resistance to fluid flow in pipes, serving as a key parameter in predicting pressure losses due to wall shear. It appears in the Darcy-Weisbach equation, which relates the pressure drop \Delta P along a pipe of length L and diameter D to the flow velocity V and fluid density \rho as follows: \Delta P = f \frac{L}{D} \frac{\rho V^2}{2} This equation captures the energy dissipation from viscous forces acting on the pipe walls, assuming fully developed, steady flow. Equivalently, the equation can be expressed in terms of head loss h_f, representing the loss in hydraulic head due to friction, where g is the acceleration due to gravity: h_f = f \frac{L}{D} \frac{V^2}{2g} This form is particularly useful in applications involving elevation changes or open-channel flows, as it directly ties friction to the effective reduction in fluid potential energy. The specifically addresses major friction losses distributed along the pipe length and does not include minor losses from fittings, valves, or expansions, which are accounted for separately using empirical loss coefficients K. The value of f is empirical, lacking a closed-form analytical solution, and depends primarily on the Reynolds number \mathrm{Re} (indicating the flow regime) and the relative roughness \varepsilon / D (where \varepsilon is the absolute roughness of the pipe interior). For laminar flows (\mathrm{Re} < 2300), f simplifies to f = 64 / \mathrm{Re}, but in turbulent regimes, it requires experimental correlations or graphical tools like the for determination. This dependence reflects the interplay between inertial forces, viscosity, and surface irregularities in real pipe systems. Historically, the equation bears the names of Henry Darcy and Julius Weisbach, who contributed foundational experimental work in the mid-19th century: Weisbach proposed an early form in 1845 based on hydraulic principles, while Darcy refined it through precise measurements in 1857, establishing its empirical basis for pipe friction. The combined nomenclature "Darcy-Weisbach" was later formalized by hydraulic engineer Hunter Rouse in the 20th century to honor both pioneers.

Chart Construction and Features

Graphical Representation

The Moody chart is constructed as a log-log plot, with the logarithm of the Reynolds number (Re) along the horizontal axis and the logarithm of the Darcy friction factor (f) along the vertical axis, facilitating the visualization of relationships across multiple orders of magnitude in pipe flow regimes. This design choice emphasizes the turbulent flow region, where most engineering applications occur, by compressing the scales appropriately without resorting to linear segments. The chart spans a Reynolds number range from $10^2 to $10^8, covering laminar, transitional, and fully turbulent flows, while the friction factor extends from $10^{-4} to 1, accommodating smooth to highly rough pipe conditions. A family of curves represents constant values of relative roughness \epsilon / D, typically ranging from $10^{-6} (near-smooth surfaces) to $10^{-1} (highly rough conduits), with each curve illustrating how f varies with Re for a fixed roughness level. Key visual elements include logarithmic grid lines for precise interpolation, prominent labeling of the smooth pipe curve (corresponding to \epsilon / D = 0) as the lower envelope of the roughness family, and a vertical line at Re = 2300 delineating the approximate onset of turbulent flow from the laminar regime. This diagram was originally published by in the 1944 issue of the ASME Transactions.

Axes and Scales

The x-axis of the Moody chart employs a base-10 logarithmic scale for the Reynolds number (Re), extending from values representing laminar flow (approximately Re = 2,000) to highly turbulent conditions (up to Re = 10^8), with major tick marks positioned at orders of magnitude to facilitate navigation across this vast range. The y-axis utilizes a similarly logarithmic scale for the Darcy friction factor (f), ranging from near-zero values typical of smooth pipes at high Re to unity in the low-Re laminar regime, enabling the depiction of f's multi-order-of-magnitude variation on a compact plot. Logarithmic scales were selected to compress the extensive dynamic ranges of both Re and f, which span several orders of magnitude in practical pipe flow scenarios, while also linearizing key empirical relationships in the turbulent regime—such as the smooth-pipe curve, which becomes an inclined straight line under this plotting method. This approach aligns with the power-law behaviors observed in turbulent friction, like the Blasius relation f ≈ 0.316 Re^{-0.25}, rendering asymptotic trends visually straightforward. The chart includes additional markings to guide interpretation, such as a dashed straight line for the laminar region defined by f = 64/Re, which follows a slope of -1 on the log-log plot, and the smooth turbulent asymptote forming the lower boundary for zero-roughness conditions. In digital adaptations, such as interactive software tools for pipe flow analysis, these logarithmic scales are preserved but enhanced with zoomable interfaces and precise interpolation capabilities to handle modern computational workflows.

Curve Families

The Moody chart consists of a family of parametric curves in the turbulent flow regime, each representing a constant relative roughness \epsilon / D, where \epsilon is the average height of surface protrusions and D is the pipe diameter. These curves, typically 15 to 20 in number, span values of \epsilon / D from 0 (smooth pipes) to 0.1, demonstrating the dependence of the Darcy friction factor f on the \mathrm{Re} for varying pipe roughness. At lower \mathrm{Re}, the curves converge, approaching similar friction factor values influenced primarily by viscous effects, while at higher \mathrm{Re}, they diverge, with rougher pipes exhibiting higher f due to increased drag from surface irregularities. The smooth pipe curve forms the lower envelope of this family and is based on Prandtl's universal law of the wall, which describes the velocity profile in the turbulent boundary layer near a smooth surface. For turbulent smooth flow, it is approximated by the implicit relation \frac{1}{\sqrt{f}} \approx 2 \log_{10} (\mathrm{Re} \sqrt{f}) - 0.8, or equivalently, f \approx \frac{1}{\left[ 2 \log_{10} (\mathrm{Re} \sqrt{f}) - 0.8 \right]^2}. This curve illustrates a monotonic decrease in f with increasing \mathrm{Re}, as the viscous sublayer thins and molecular viscosity plays a reduced role in momentum transfer. In the fully rough regime, prevalent at high \mathrm{Re}, the curves for nonzero \epsilon / D approach horizontal asymptotes where f becomes independent of \mathrm{Re} and depends only on relative roughness. This behavior is captured by the approximation f \approx \frac{1}{\left[ 2 \log_{10} \left( \frac{3.7}{\epsilon / D} \right) \right]^2}, derived from Nikuradse's experiments on artificially roughened pipes, reflecting the dominance of form drag from roughness elements that extend beyond the viscous sublayer. Each curve transitions through three zones: the hydraulically smooth zone at moderate \mathrm{Re}, where roughness is submerged in the viscous sublayer and f follows the smooth pipe curve; the transitional zone at intermediate \mathrm{Re}, where roughness partially disrupts the sublayer, causing f to deviate upward from the smooth curve; and the fully rough zone at high \mathrm{Re}, marked by the Re-independent asymptotes. These zones underscore the shift from viscosity-dominated to roughness-dominated resistance in turbulent pipe flow. For \epsilon / D values not coinciding with the chart's predefined curves, engineers often apply linear interpolation between adjacent curves, typically on a logarithmic scale for f versus \mathrm{Re}, to obtain an approximate friction factor value suitable for design calculations.

Interpretation and Usage

Determining Friction Factor

To determine the Darcy friction factor using the Moody chart, the process begins with calculating the Reynolds number (Re) and relative roughness (ε/D) for the pipe flow scenario. The Reynolds number is computed from fluid properties such as density and viscosity, along with pipe diameter and flow velocity derived from the volumetric flow rate. Relative roughness is the ratio of the pipe's absolute roughness height (ε) to its internal diameter (D), with ε values typically sourced from material-specific tables for new or aged pipes. The core procedure involves locating these parameters on the chart. First, identify the Re value on the horizontal x-axis, which spans logarithmic scales from laminar to highly turbulent regimes. Then, identify the curve corresponding to the calculated ε/D. Draw a vertical line upward from the Re point to intersect the selected ε/D curve, and from that intersection, draw a horizontal line to the y-axis to read the friction factor (f). This yields f directly for fully turbulent flow (Re > 4000). If the intersection falls in the region (approximately 2300 < Re < 4000), where curves are dashed to indicate instability, an initial f reading may require iteration: use the obtained f to refine velocity or Re calculations if flow rate is unknown, and repeat until convergence. Essential inputs for this process include known fluid properties (density, viscosity, temperature-dependent), pipe dimensions (diameter, roughness), and flow rate or velocity. Potential error sources encompass inaccuracies in interpolating between curve families on printed charts, limitations in chart resolution for very low or high Re values, and misestimation of ε for fouled pipes, which can alter effective roughness by factors of 2 to 4. For verification, particularly when Re < 2300 indicates laminar flow, cross-check the chart-derived f against the analytical Hagen-Poiseuille relation, f = 64/Re, as the Moody chart's laminar line follows this exactly but may be less precise at low Re due to graphical scaling. Modern digital tools, such as engineering software incorporating digitized Moody charts, automate this procedure for improved accuracy and iteration handling without manual reading errors.

Transition and Laminar Regions

In the laminar region of the Moody chart, applicable for Reynolds numbers Re < 2300, the Darcy friction factor follows the exact relation f = \frac{64}{\text{Re}}, derived from the governing fully developed, steady, incompressible flow of Newtonian fluids in circular pipes. This linear relationship on a log-log scale appears as a straight line with a slope of -1 and is independent of relative roughness, as viscous shear dominates the momentum transfer without significant influence from wall irregularities. The , originally established through experimental and theoretical work in the mid-19th century, provides precise head loss predictions under these conditions without reliance on empirical correlations. The critical Reynolds number marking the onset of instability in pipe flow is approximately 2300, below which laminar conditions prevail and above which turbulence may emerge, as identified in Osborne Reynolds' foundational 1883 experiments using dye injection to visualize flow regimes in tubes. This threshold represents the point where inertial forces begin to overcome viscous damping, leading to potential sinuous motion and flow breakdown. However, the exact value can vary slightly based on experimental conditions, with Reynolds reporting values around 2000 under controlled settings. The transition zone, spanning 2300 < Re < 4000, exhibits unstable behavior with intermittent laminar and turbulent patches, resulting in considerable scatter of friction factor data on the Moody chart, often depicted as a shaded or hatched band to indicate uncertainty. In this regime, the friction factor cannot be precisely determined from the chart alone due to the flow's sensitivity to perturbations, and conservative estimates—typically selecting higher values within the scatter band—are recommended for engineering designs to account for potential increases in pressure drop. Factors such as entrance geometry and upstream disturbances can accelerate transition by introducing vorticity or unsteadiness, effectively lowering the critical Reynolds number and exacerbating unpredictability. Overall, the Moody chart's utility diminishes in the laminar and transition regions compared to fully turbulent flows, where empirical curves are more reliable; for laminar cases, direct application of the Hagen-Poiseuille-derived formula is preferred to avoid graphical inaccuracies.

Example Calculations

To illustrate the practical use of the Moody chart in determining the Darcy friction factor and subsequent pressure losses, consider water flowing through a horizontal pipe with diameter D = 0.1 m and length L = 100 m at 20°C, where the dynamic viscosity is \mu = 1.00 \times 10^{-3} Pa·s and density is \rho = 1000 kg/m³. The flow rate Q is adjusted for each example to achieve the specified Reynolds number, with velocity V = \frac{4Q}{\pi D^2} and \text{Re} = \frac{\rho V D}{\mu}. The pressure drop is calculated using the Darcy-Weisbach equation \Delta P = f \frac{L}{D} \frac{\rho V^2}{2}, where f is the friction factor. Example 1: Laminar flow in a smooth tube (Re = 1000)
For laminar flow in a smooth pipe, the friction factor is given exactly by f = \frac{64}{\text{Re}}. With Re = 1000, f = \frac{64}{1000} = 0.064. This corresponds to V = \frac{\text{Re} \mu}{\rho D} = 0.01 m/s and Q \approx 7.85 \times 10^{-5} m³/s. The pressure drop is \Delta P = 0.064 \times \frac{100}{0.1} \times \frac{1000 \times (0.01)^2}{2} = 3.2 Pa, or head loss h_f = \frac{\Delta P}{\rho g} \approx 3.3 \times 10^{-4} m (with g = 9.81 m/s²), verifying the low-loss regime for slow viscous flows.
Example 2: Turbulent flow in a rough steel pipe (Re = $10^5, \epsilon / D = 0.00015)
Locate Re = $10^5 on the Moody chart and follow the curve for relative roughness \epsilon / D = 0.00015 (typical for drawn tubing approximating new steel); interpolation yields f \approx 0.02 in the transitionally rough turbulent region. This gives V = 1 m/s and Q \approx 0.00785 m³/s. The pressure drop is \Delta P = 0.02 \times 1000 \times \frac{1000 \times 1^2}{2} = 10,000 Pa (10 kPa), or h_f \approx 1.02 m, highlighting moderate losses due to combined viscous and roughness effects.
Example 3: High-Re fully rough regime (Re = $10^7, \epsilon / D = 0.01)
In the fully rough turbulent regime at high Re, f becomes independent of Re and depends only on \epsilon / D; from the Moody chart horizontal asymptote for \epsilon / D = 0.01 (representative of highly corroded pipes), f \approx 0.038. Here, V = 100 m/s and Q \approx 0.785 m³/s. The pressure drop is \Delta P = 0.038 \times 1000 \times \frac{1000 \times 100^2}{2} = 1.9 \times 10^8 Pa (190 MPa), or h_f \approx 19,400 m, demonstrating severe losses dominated by surface roughness.
The following table summarizes the key numerical results for comparison across regimes:
ExampleRe\epsilon / DfV (m/s)\Delta P (Pa)h_f (m)
1 (Laminar)100000.0640.013.20.00033
2 (Turbulent)$10^50.000150.02110,0001.02
3 (Fully Rough)$10^70.010.038100$1.9 \times 10^819,400

Historical Development

Origins in Early 20th Century

The foundations of friction factor correlations trace back to mid-19th-century hydraulic experiments that established the basic proportionality of head loss in pipe flow. In 1845, Julius Weisbach formulated an empirical equation for frictional head loss, expressing it as proportional to the pipe length, inversely proportional to the diameter, and to the square of the average velocity, introducing a dimensionless to account for viscous effects. This work built on earlier observations of laminar flow resistance by Gotthilf Hagen and in the 1830s and 1840s. A decade later, in 1857, Henry Darcy conducted systematic experiments on water flow through pipes of varying diameters and lengths, confirming Weisbach's velocity-squared proportionality and refining the through direct measurements of pressure drops, which highlighted its dependence on pipe geometry and flow conditions. These pioneering studies provided the empirical basis for quantifying frictional resistance but lacked a unified framework for the coefficient across diverse flow regimes. By the early 20th century, efforts to visualize friction factor variations emerged through graphical representations. In 1914, Thomas Ernest Stanton and J. R. Pannell published an influential diagram plotting the friction factor against the Reynolds number for flows in smooth pipes and those with varying roughness, based on experimental data from multiple pipe materials and fluids. This Stanton-Pannell chart, one of the earliest such plots, illustrated the decline in friction factor with increasing Reynolds number in the turbulent regime for smooth walls, offering engineers a practical tool to interpolate values from scattered test results without relying solely on algebraic formulas. It emphasized the role of the Reynolds number in distinguishing laminar and turbulent behaviors but was limited to specific pipe types and did not fully integrate roughness effects. Theoretical advancements in the 1930s further refined understanding of friction in smooth pipes through boundary layer concepts. Ludwig Prandtl, building on his 1904 boundary layer theory, developed universal velocity profiles for turbulent wall-bounded flows using mixing-length hypotheses, which described the logarithmic distribution of velocity near the wall. These profiles, detailed in Prandtl's 1932 and 1933 works, enabled derivations of friction factor correlations for smooth pipes, such as expressions linking the factor inversely to the logarithm of the , capturing the asymptotic behavior in fully turbulent flow without roughness influences. This conceptual framework shifted focus from purely empirical fits to physically grounded models of near-wall turbulence. Experimental quantification of roughness effects came prominently from Johann Nikuradse's controlled tests in the 1930s. In his 1933 study, Nikuradse coated commercial pipes with uniform sand grains of controlled sizes to simulate idealized roughness, measuring pressure drops across a wide range of Reynolds numbers in turbulent flows. These experiments systematically demonstrated how the relative roughness (ratio of grain height to pipe diameter, ε/D) shifts the friction factor upward from smooth-pipe values, particularly in fully rough regimes where it becomes independent of Reynolds number, establishing ε/D as a key parameter for turbulent friction. Nikuradse's data provided a benchmark for distinguishing hydraulically smooth, transitionally rough, and fully rough conditions. Despite these advances, friction factor determination remained challenging due to fragmented empirical data spanning diverse pipe materials, fluids, and experimental setups, with correlations often limited to narrow parameter ranges or requiring iterative solutions. Engineers relied on disparate tables, charts, and formulas—like those from in 1930 compiling —which lacked a comprehensive synthesis for practical application across all regimes. This proliferation of incomplete models underscored the need for a unified graphical tool to consolidate the growing body of experimental and theoretical insights. These early developments culminated in the graphical unification presented in .

Lewis F. Moody's Contribution

Lewis Ferry Moody (1880–1953) was an American mechanical engineer and pioneering academic in hydraulic engineering, best known for his contributions to fluid mechanics and turbomachinery. He served as the first Professor of Hydraulics in Princeton University's School of Engineering, where he also taught fluid mechanics and machine design, and held more than 90 patents for innovations in hydraulic turbines. Moody's career bridged academia and industry, with practical experience in hydraulic systems that informed his research on pipe flow dynamics. In 1944, Moody published the seminal paper "Friction Factors for Pipe Flow" in the Transactions of the American Society of Mechanical Engineers (Vol. 66, pp. 671–678), aiming to provide engineers with a straightforward method for estimating friction factors in clean, new pipes to calculate head loss. Building briefly on early 20th-century experimental foundations, Moody compiled and plotted data from key sources, including Johann Nikuradse's 1933 experiments on artificially roughened pipes, C. F. Colebrook's 1938–1939 implicit equation for turbulent flow transitions, and earlier correlations for laminar and smooth turbulent regimes by researchers like Blasius and Prandtl. Moody's key innovation was a single, unified logarithmic diagram plotting the Darcy-Weisbach friction factor against the Reynolds number for a range of relative roughness values, seamlessly integrating laminar flow, smooth turbulent flow, and fully rough turbulent regimes into one accessible graph. He also addressed notation inconsistencies by noting the relationship between the Darcy friction factor and the Fanning friction factor, where the latter is one-fourth of the former, facilitating its use across different engineering conventions. This graphical approach simplified complex calculations that previously required separate equations or tables, making it practical for design applications in piping systems. The publication had a profound impact on post-World War II engineering practice, standardizing friction factor estimation in industries like chemical processing, power generation, and water supply, where accurate pipe flow predictions were essential for efficient system design. Moody's chart was rapidly adopted and reprinted in authoritative handbooks, including Marks' Standard Handbook for Mechanical Engineers and Perry's Chemical Engineers' Handbook, ensuring its enduring utility. Its legacy persists as the "Moody chart," with minor updates in later editions to refine curve fits based on additional data, though the original framework remains foundational in fluid engineering education and practice.

Fanning Friction Factor

The Fanning friction factor, denoted as f_F, is a dimensionless parameter defined as the ratio of the wall shear stress \tau_w to the dynamic pressure of the fluid, expressed as f_F = \frac{\tau_w}{\rho V^2 / 2}, where \rho is the fluid density and V is the average flow velocity. This definition directly ties the factor to local shear forces at the conduit wall, making it particularly suited for analyses involving momentum transfer and boundary layer effects. It is related to the Darcy friction factor f_D by the scaling f_F = f_D / 4, a convention that arises because the Darcy formulation incorporates an additional factor of 4 in the pressure drop expression for pipe flow. In the context of the , which plots the Darcy friction factor against the Reynolds number and relative roughness, values for the Fanning factor can be obtained by dividing the chart's y-axis readings by 4. Moody's original 1944 chart used Darcy notation, but the direct scaling allows adaptation for Fanning applications without altering the underlying correlations. The Fanning friction factor finds prominent use in chemical engineering, especially for calculating pressure drops in annular ducts and heat exchangers, where shear stress distributions are critical for design. It appears frequently in older engineering literature for both open channel flows and pipe analogies, reflecting its origins in early hydraulic studies. Named after American hydraulic engineer (1837–1911), the factor was introduced in his 1877 treatise A Practical Treatise on Hydraulic and Water-Supply Engineering, where he compiled empirical resistance data and adapted earlier formulations for practical water supply systems. For general pipe flow calculations, the Darcy friction factor is preferred due to its standardization in the for head loss. The Fanning factor, however, is more appropriate for boundary layer analyses and momentum balance problems, such as those in non-circular ducts or where wall shear is the primary concern.

Colebrook-White Equation

The provides an implicit relationship for calculating the Darcy friction factor f in turbulent pipe flow, incorporating both the effects of surface roughness and viscous forces through the \mathrm{Re} and relative roughness \epsilon/D. The equation is given by \frac{1}{\sqrt{f}} = -2 \log_{10} \left( \frac{\epsilon}{3.7D} + \frac{2.51}{\mathrm{Re} \sqrt{f}} \right), where the first term in the logarithm represents the rough pipe contribution derived from the fully rough regime, and the second term accounts for the smooth pipe behavior in the transitional turbulent flow. This form unifies the smooth and rough turbulent regimes, making it applicable across a wide range of pipe conditions. The derivation of the Colebrook-White equation stems from the Prandtl-von Kármán law of the wall, which describes the universal velocity profile in the turbulent boundary layer near the pipe wall. By matching the logarithmic velocity profiles in the overlap region between the inner wall layer and the outer defect layer, and incorporating empirical adjustments based on experimental data from artificially roughened pipes, Colebrook blended the smooth-wall Prandtl equation with the fully rough von Kármán relation using a logarithmic interpolation technique. This approach ensures a smooth transition between regimes without arbitrary switching functions. Solving the Colebrook-White equation requires iterative numerical methods due to its transcendental nature, with the Newton-Raphson method being a common choice for converging on f from an initial guess, such as the Blasius approximation for smooth pipes. Explicit approximations, like the Haaland equation, offer direct solutions with minimal error; for instance, Haaland's formula rearranges the logarithmic terms to yield f explicitly as \frac{1}{\sqrt{f}} \approx -1.8 \log_{10} \left[ \left( \frac{\epsilon/D}{3.7} \right)^{1.11} + \frac{6.9}{\mathrm{Re}} \right], providing accuracy comparable to the implicit form for practical engineering use. These methods are essential for computational efficiency in design applications. The equation demonstrates high fidelity to experimental data, fitting the measurements of Nikuradse on rough pipes within approximately 1% across the turbulent regime for \mathrm{Re} > 4000, though accuracy diminishes slightly in the critical near \mathrm{Re} \approx 2300. It is considered the standard for turbulent prediction in circular pipes under steady, conditions. The Moody chart graphically represents the numerical solutions to the Colebrook-White equation, plotting iso-curves of constant relative roughness \epsilon/D against \mathrm{Re} to visualize how f varies, thereby approximating the implicit model's outputs for manual interpolation in engineering practice.

Modern Computational Alternatives

With the advent of computational tools, explicit approximations to the Colebrook-White equation have emerged as direct alternatives to graphical interpolation on the Moody chart, enabling rapid and accurate calculations without iteration. One prominent example is the Swamee-Jain equation, which provides an explicit formula for the Darcy f in turbulent : f = \frac{0.25}{\left[ \log_{10} \left( \frac{\epsilon}{3.7D} + \frac{5.74}{\mathrm{Re}^{0.9}} \right) \right]^2} where \epsilon is the absolute roughness, D is the pipe diameter, and \mathrm{Re} is the Reynolds number; this approximation yields errors less than 1% across a wide range of relative roughnesses (\epsilon/D from $10^{-6} to $10^{-2}) and Reynolds numbers (from 5000 to $10^8). Beyond explicit formulas, modern software implementations have further supplanted manual chart usage by solving the implicit Colebrook-White equation numerically or integrating full fluid dynamics simulations. Computational fluid dynamics (CFD) codes, such as ANSYS Fluent, compute friction factors as part of broader simulations for pipe networks and complex geometries, incorporating wall shear stress directly from Navier-Stokes solutions. Spreadsheet tools like Microsoft Excel with built-in solvers facilitate iterative solutions for straightforward pipe flow problems, while Python libraries, including SciPy's fsolve function, allow programmatic resolution of the Colebrook equation in custom scripts for batch processing or integration into larger engineering models. These computational alternatives offer significant advantages over the Moody chart, including higher precision (limited only by numerical ), faster computation for studies, and the ability to handle non-ideal conditions such as irregular geometries, transient flows, and multiphase interactions that charts cannot represent. They also enable seamless integration with optimization algorithms for design refinement, reducing trial-and-error in applications like sizing. Despite these advances, the Moody chart persists in scenarios requiring quick hand calculations, educational demonstrations of trends, or preliminary estimates where computational resources are unavailable. Extended computational models build on chart principles to address limitations in specialized flows, incorporating non-Newtonian through generalized friction correlations or microscale effects via modified roughness scaling in microchannel simulations.

References

  1. [1]
    Internal Flows – Introduction to Aerospace Flight Vehicles
    A standard Moody chart for estimating the Darcy-Weisbach friction factor in pipes and ducts. ... Reynolds number and relative roughness. Ensure that the Reynolds ...
  2. [2]
    Friction Factors for Pipe Flow | J. Fluids Eng. - ASME Digital Collection
    Dec 14, 2022 · The object of this paper is to furnish the engineer with a simple means of estimating the friction factors to be used in computing the loss of head in clean ...
  3. [3]
    None
    - **Definition**: Reynolds number (Re) is a dimensionless group indicating flow type, named after Osborne Reynolds, who studied laminar-to-turbulent transition in 1883.
  4. [4]
    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 ...Missing: physical | Show results with:physical
  5. [5]
    [PDF] Nondimensionalization of the Navier-Stokes Equation
    Nondimensionalization: We begin with the differential equation for conservation of linear momentum for a. Newtonian fluid, i.e., the Navier-Stokes equation. For ...Missing: pipe | Show results with:pipe
  6. [6]
    [PDF] Chapter 15: Turbulence [version 1215.1.K] - Caltech PMA
    We also defined the Reynolds number; for pipe flow it is Red ≡ ¯vd/ν, where ¯v is the mean speed in the pipe. Now, it turns out experimentally that the ...
  7. [7]
    [PDF] Reynolds Number Worksheet - LabSmith
    Since the Reynolds Number is a dimensionless number, all units (m, kg, etc.) will cancel out for the final value. For water at ~20 C. 𝜌. = 1000 kg/m3. 𝜇. = 1.0 ...
  8. [8]
    [PDF] Pipe Flows - Purdue Engineering
    Dec 15, 2021 · The Darcy friction factor appears in the Moody chart for incompressible, viscous pipe flow. Note again that this solution is only valid only ...
  9. [9]
    Head Loss | Engineering Library
    The quantity used to measure the roughness of the pipe is called the relative roughness, which equals the average height of surface irregularities (ε) divided ...
  10. [10]
    Relative Roughness of Pipe | Calculation | nuclear-power.com
    Relative roughness is the average height of surface irregularities divided by the pipe diameter, used to measure the roughness of a pipe's inner surface.
  11. [11]
    Absolute Roughness of Pipe Material - Neutrium
    May 19, 2012 · Absolute Roughness for Common Materials ; Drawn Tubing, Glass, Plastic, 0.0015-0.01 ; Drawn Brass, Copper, Stainless Steel (New), >0.0015-0.01.
  12. [12]
    Pipe Roughness - Pipe Flow Software
    The relative roughness of a pipe is its roughness divided by its internal diameter or e/D, and this value is used in the calculation of the pipe friction ...Missing: engineering | Show results with:engineering
  13. [13]
    Surface Roughness Coefficients - The Engineering ToolBox
    Relative Roughness. Relative roughness - the ratio between absolute roughness an pipe or duct diameter - is important when calculating pressure loss in ducts ...Missing: definition | Show results with:definition
  14. [14]
    ASME B46.1-2019: Surface Texture (Roughness, Waviness, Lay)
    Aug 7, 2020 · Breakdown of ASME B46.1-2019 sections and its changes to cover surface texture and its constituents: roughness, waviness, and lay.
  15. [15]
    Absolute Pipe Roughness - EnggCyclopedia
    Following table gives typical roughness values in millimeters for commonly used piping materials. Surface Material, Absolute Roughness Coefficient - ε in mm.
  16. [16]
    Laws of Flow in Rough Pipes
    An experimental investigation is made of the turbulent flow of water in pipes with various degrees of relative roughness.
  17. [17]
    Roughness effects in turbulent pipe flow | Journal of Fluid Mechanics
    Sep 15, 2006 · The flow exhibits hydraulically smooth, transitionally rough, and fully rough behaviours. The surface of the pipe was prepared with a honing tool.
  18. [18]
    [PDF] Head Loss in Pipe Systems Laminar Flow and Introduction to ...
    Jan 23, 2007 · For turbulent flow the dimensional form of the equation for pressure drop is ∆p = φ(V, L, D, µ, ρ, ε) where ε is the length scale determining ...
  19. [19]
    [PDF] Chapter 13: Head Loss in Pipes - eCommons - University of Dayton
    Dec 22, 2022 · For laminar flow the Darcy-Weisbach friction factor is only a function of the Reynolds Number as illustrated below. Interestingly, notice that ...
  20. [20]
    USBR Water Measurement Manual - Chapter 2
    The Darcy-Weisbach friction factor, f, is nondimensional and is a function of Reynolds number, 4RhV/L, and relative roughness, k/4Rh, in which L is kinematic ...Missing: formula | Show results with:formula
  21. [21]
    [PDF] A correlation of formulas for the flow of fluids in pipes - Scholars' Mine
    In about the middle of the nineteenth century 2, much used pipe formula came into use. (2 ) Credit for its origin is given to Darcy, Weisbach, rannin8, or ...
  22. [22]
    [PDF] Technical Note: Friction Factor Diagrams for Pipe Flow
    Oct 2, 2011 · This technical note describes diagrams of friction factor for pipe flow that have been prepared using, mainly, the equations that Lewis Moody.
  23. [23]
    [PDF] Friction Factors for Pipe Flow
    BY LEWIS F. MOODY,' PRINCETON, N. J.. The object of this paper is to furnish the engineer with a simple means of estimating the friction ...
  24. [24]
    Moody chart (diagram) | tec-science
    Jul 4, 2020 · The Moody diagram is a chart showing the Darcy friction factor of a pipe as a function of the Reynolds number for selected roughnesses of the pipe wall.
  25. [25]
    Friction Factor Calculations - Pipe Flow Software
    The friction factor or Moody chart is the plot of the relative roughness (e/D) of a pipe against the Reynold's number. The blue lines plot the friction factor ...
  26. [26]
    [PDF] Some Fully-Constrained Turbulent Flows
    Schlichting (1968) suggested that "Prandtl's. Universal Law of Friction for Smooth Pipes". 庐. = 2.0 log | Re√F-. -0.8. (5.11) fits the data shown in Figure 5.9 ...
  27. [27]
    [PDF] Rough Pipe Flows
    Figure 1: The Moody diagram, the friction factor, f, for flow in a circular pipe plotted against the Reynolds number, 2V R/ν.Missing: chart | Show results with:chart
  28. [28]
    [PDF] The Moody Chart
    Moody. (1880–1953) redrew Rouse's diagram into the form commonly used today. The now famous Moody chart is given in the appendix as Fig. A–12. It presents the ...
  29. [29]
    Reading the Moody chart with a linear interpolation method - Nature
    Apr 21, 2022 · (a) A possible error when reading the Moody chart. The relative roughness ε/ds of the lines from the top to the bottom are 0.05, 0.04, 0.03, 0. ...
  30. [30]
    Moody's Friction Factor - an overview | ScienceDirect Topics
    The Moody friction factor can be obtained by plotting the Darcy friction factor as a function of Reynolds number and relative roughness.
  31. [31]
  32. [32]
  33. [33]
  34. [34]
    [PDF] The History of the Darcy-Weisbach Equation for Pipe Flow Resistance
    A concise examination of the evolution of the equation itself and the Darcy friction factor is presented from their inception to the present day. The.<|control11|><|separator|>
  35. [35]
    [PDF] Reynolds_1883.pdf
    On one end of the tube being slightly raised the water would flow to the upper end and the bisulphide fall to the lower, causing opposite currents along the ...
  36. [36]
    [PDF] Absolute viscosity of water at 20-degrees-C
    The absolute viscosity of water at 20°C is 0.010019 poise, with a range of 0.010019 ± 0.0003.<|control11|><|separator|>
  37. [37]
    Pure Water Density Standard, UKAS ISO/IEC17025 and ... - PubChem
    Pure Water Density Standard, UKAS ISO/IEC17025 and ISO Guide 34 certified, density: 0.9982 g/mL at 20 C, density: 0.9970 g/mL at 25 C.
  38. [38]
    [PDF] Friction Factors for Pipe Flow
    BY LEWIS F. MOODY,' PRINCETON, N. J. The object of this paper is to furnish the engineer with a simple means of estimating the friction factors to be used in ...
  39. [39]
    Henry Darcy and the making of a law - Brown - 2002 - AGU Journals
    Jul 17, 2002 · Darcy proposed an equation similar to (4) with friction coefficients that were functions of D, and showed it reduced to a dimensionally ...
  40. [40]
    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 ...Missing: factor | Show results with:factor
  41. [41]
    Lewis Ferry Moody - EPFL Graph Search
    Lewis Ferry Moody (5 January 1880 – 21 February 1953) was an American engineer and professor, best known for the Moody chart, a diagram capturing ...Missing: biography chemical
  42. [42]
    Lewis Moody - Prabook
    Lewis Moody. engineer. Lewis Ferry Moody was an American engineer and professor, best known for the Moody chart, a diagram capturing relationships between ...Missing: biography | Show results with:biography
  43. [43]
    Obituary for Lewis Ferry Moody - Newspapers.com™
    90 Patents Funeral services for Lewis Ferry Moody, professor emeritus of hydraulic engineering at Princeton University and inventor of numerous improvements in ...
  44. [44]
    Fanning Friction Factor - an overview | ScienceDirect Topics
    The Moody friction factor can be obtained by plotting the Darcy friction factor as a function of Reynolds number and relative roughness. The Moody friction ...
  45. [45]
    Fanning friction factor - A Website For Engineers - Process Associates
    Fanning friction factor is proportional to shear stress at pipe/conduit wall as number of velocity heads and is used in momentum transfer in general and ...
  46. [46]
    FRICTION FACTORS FOR SINGLE PHASE FLOW IN SMOOTH AND ...
    The friction factor of pipes will be the same of their Reynolds number, roughness patterns, and relative roughness are the same.<|control11|><|separator|>
  47. [47]
    Pipe flow: a gateway to turbulence | Archive for History of Exact ...
    Oct 2, 2020 · By the 1860s, the experiments of hydraulic engineers like Hagen, Weisbach, Darcy and others had abundantly made clear that the flow in very ...
  48. [48]
    19.3 Frictional Pipe Flow | Process Fluid Mechanics - InformIT
    Jul 18, 2018 · ... Fanning friction factor. (The Fanning friction factor is typically used by chemical engineers. There is also the Moody friction factor ...
  49. [49]
    [PDF] Laminar Flow Forced Convection Heat Transfer and Flow Friction in ...
    Jan 21, 1972 · "Fanning" or "small" friction factor, for fully developed flow if no ... for the annular ducts with axial exponential heat flux dis-.
  50. [50]
    Explicit Equations for Pipe-Flow Problems | Vol 102, No 5
    Feb 3, 2021 · Reported herein are explicit and accurate equations for pipe diameter and head loss and a closed form solution for the discharge through the pipe.
  51. [51]
    [PDF] Moopr,LING THE TURBULENT FLow oF NoN-NnwtoNIAN
    Figure 8 shows the effect of differept particle sizes for the same rheology. This is analogous to the relative roughness effect shown on the Moody diagram.<|control11|><|separator|>
  52. [52]
    [PDF] Effects of structured roughness on fluid flow at the microscale level
    Nov 19, 2010 · This constricted flow approach led to a modified. Moody diagram which covers higher relative roughness values and smaller hydraulic diameters.Missing: chart families