Fact-checked by Grok 2 weeks ago

Fanning friction factor

The Fanning friction factor, denoted as f_f, is a in that characterizes the frictional resistance to fluid flow in , ducts, and conduits by relating wall to the fluid's . It is defined as f_f = \frac{\tau_w}{\rho V^2 / 2}, where \tau_w is the wall , \rho is the fluid density, and V is the average . Equivalently, it connects to via f_f = \frac{D_h \Delta P}{2 L \rho V^2}, with D_h as the , \Delta P as the pressure loss over length L. Named after American hydraulic engineer John Thomas Fanning (1837–1911), who introduced tabulated values based on earlier experiments by Henry Darcy in his 1877 treatise A Practical Treatise on Hydraulic and Water-Supply Engineering, the factor has become a standard tool for predicting energy dissipation due to viscosity and surface roughness. Unlike the Darcy-Weisbach friction factor f_d, which is four times larger (f_d = 4 f_f), the Fanning version emphasizes local shear effects and is prevalent in chemical and mechanical engineering applications for non-circular conduits using the hydraulic diameter. For laminar flow in smooth pipes, it simplifies to f_f = \frac{16}{\text{Re}}, where Re is the Reynolds number, while turbulent regimes require empirical correlations like the Colebrook equation: \frac{1}{\sqrt{f_f}} = -4 \log_{10} \left( \frac{\epsilon / D_h}{3.7} + \frac{1.255}{\text{Re} \sqrt{f_f}} \right), accounting for relative roughness \epsilon / D_h. This factor underpins calculations for , pump sizing, and design, enabling efficient system optimization across industries like processing and HVAC. Its use in the balance equation, \frac{\Delta P}{\rho} + \frac{\Delta V^2}{2} + g \Delta z = -2 f_f \frac{L}{D_h} V^2, integrates with other losses for comprehensive flow analysis.

Introduction

Definition and Physical Meaning

The Fanning friction factor, denoted as f, is a dimensionless quantity that characterizes the frictional losses in fluid flow through conduits, particularly in pipes and ducts. It is defined as the ratio of the wall shear stress \tau_w to the dynamic pressure of the flow, expressed mathematically as f = \frac{\tau_w}{\rho u_m^2 / 2}, where \rho is the fluid density and u_m is the mean flow velocity. This definition arises in the context of internal flows where viscous effects at the solid boundary dominate the energy dissipation. Physically, the Fanning friction factor represents the proportion of the flow's that is dissipated as due to at the wall per unit volume. The wall \tau_w quantifies the tangential force exerted by the on the conduit surface, while \rho u_m^2 / 2 is the kinetic energy density () associated with the bulk motion of the . Thus, f provides a measure of how effectively the flow's is retarded by boundary , influencing the overall loss and energy requirements in systems such as pipelines and heat exchangers. The derivation of the Fanning friction factor stems from the momentum balance for steady, fully developed flow, which is an integrated form of the Navier-Stokes equations applied to a within a circular . Considering a cylindrical section of length L and diameter D, the axial pressure force \pi (D/2)^2 \Delta P balances the frictional \tau_w \cdot \pi D L acting on the wall, yielding the relation \frac{\Delta P}{L} = \frac{4 \tau_w}{D}. Substituting the definition of f gives the equivalent form f = \frac{\Delta P \, D}{2 \rho u_m^2 L}, which directly links the observable to the flow parameters and friction. This formulation assumes incompressible, behavior and neglects entrance effects or secondary flows. The applicability of the Fanning friction factor depends on the flow regime, which is determined by the \mathrm{Re}, a dimensionless group indicating the relative importance of inertial to viscous forces in the flow.

History and Development

The Fanning friction factor is named after John Thomas Fanning (1837–1911), an American hydraulic engineer who introduced it in his 1877 publication, A Practical Treatise on Water-supply Engineering, initially for applications in and systems. Fanning compiled empirical tables of friction coefficients from diverse experimental data sources, including those from French, American, English, and German studies, to quantify pressure losses in conduits. The concept's early development traces its roots to the Darcy-Weisbach equation, formulated in 1857 by and Julius Weisbach, which expressed head loss as a function of , length, and a dependent on and roughness. Fanning adapted this framework in the late for broader engineering contexts, particularly emphasizing hydraulic radius over , which resulted in his being one-fourth the value of the Darcy variant; this adaptation gained traction in for handling non-circular ducts and process . A key milestone occurred with its formal adoption in Perry's Chemical Engineers' Handbook in its first edition of 1934, where it became a standard tool for fluid flow calculations in . The distinction between the Fanning and Darcy friction factors solidified in mid-20th-century engineering texts, as literature consistently favored the Fanning form for momentum transfer analyses, while civil and increasingly aligned with the version in hydraulic designs. The evolution of the Fanning friction factor shifted post-1900 from reliance on empirical tables to dimensionless correlations, enabling predictions across scales via parameters like the . This transition was significantly influenced by Johann Nikuradse's 1933 experiments on artificially roughened pipes, which provided foundational data on and roughness effects, later integrated into universal charts.

Relation to Other Friction Factors

Comparison with Darcy Friction Factor

The Darcy friction factor, denoted as f_D, and the Fanning friction factor, denoted as f_F, are related by the equation f_D = 4 f_F. Both factors quantify the frictional resistance in fluid flow through pipes, describing the same underlying but normalized differently: the Fanning factor directly relates to the wall normalized by , while the Darcy factor incorporates an additional factor of four to align with head loss conventions in certain applications. The pressure drop \Delta P across a pipe of length L and diameter D due to friction can be expressed using either factor. In the Darcy-Weisbach form, the equation is \Delta P = f_D \frac{L}{D} \frac{\rho u_m^2}{2}, where \rho is the fluid density and u_m is the mean velocity. Equivalently, in the Fanning form, it is \Delta P = 4 f_F \frac{L}{D} \frac{\rho u_m^2}{2}. These variants ensure consistency in calculations despite the differing definitions. Historically, the divergence between the two arose from disciplinary preferences in engineering practice. The Darcy friction factor emerged in civil and contexts, emphasizing head loss calculations for water conveyance and large-scale systems, building on the work of and Julius Weisbach in the mid-19th century. In contrast, the Fanning friction factor, introduced by John Thomas Fanning in 1877 through compilations of experimental data, gained prominence in chemical and , where focus on wall facilitated analyses of and multiphase flows in smaller conduits. In graphical representations like the Moody diagram, which plots friction factor against and relative roughness, the values correspond to the Darcy friction factor; thus, Fanning friction factor values read from such charts must be divided by four to obtain the appropriate f_F. This conversion is essential for cross-disciplinary applications to avoid errors in predictions.

Usage in Engineering Contexts

The Fanning friction factor is particularly favored in chemical and disciplines due to its direct linkage with heat and analogies, such as the Chilton-Colburn , where the Colburn j-factor for is expressed as j_H = \frac{f}{2}, with f denoting the Fanning friction factor. This relation facilitates the between transfer () and transfer in convective processes, enabling engineers to predict performance and efficiencies without separate empirical correlations. In these fields, the factor's emphasis on local wall aligns with analyses common in and equipment sizing. The Fanning friction factor integrates seamlessly with the , defined as \text{Re} = \frac{\rho u_m D}{\mu}, to delineate flow regimes and select appropriate correlations for friction losses. This dimensionless pairing allows practitioners to transition between laminar and turbulent regimes systematically, with f = \frac{16}{\text{Re}} governing laminar flows and more complex functions for turbulent ones, providing a unified framework for fluid dynamic assessments in and duct systems. In standards and tools, the Fanning friction factor appears prominently in references like , where it is employed for estimations in process industries, and in ASME codes such as the Boiler and Pressure Vessel Code, which incorporate it for safe design of fluid-handling components. For visualization, the —originally plotted for the friction factor—can be adapted for Fanning usage by scaling the y-axis values by \frac{1}{4}, since the Fanning factor is one-quarter of the value, simplifying interpolation for relative roughness and effects. Its advantages shine in handling multiphase flows and non-Newtonian fluids, where the direct to wall shear stress, \tau_w = f \frac{\rho u_m^2}{2}, simplifies local stress computations without additional geometric factors inherent in head loss formulations. This local focus proves invaluable for modeling slurry transport or polymer processing, reducing complexity in rheological adjustments compared to global metrics.

Formulas for Laminar Flow

Circular Tubes

In fully developed through circular tubes, the Fanning friction factor is given by the formula f = \frac{16}{\mathrm{Re}} where \mathrm{Re} is the , defined as \mathrm{Re} = \frac{\rho V D}{\mu} with \rho as fluid density, V as average velocity, D as tube diameter, and \mu as dynamic viscosity. This relation is derived from the Hagen-Poiseuille law, which provides an exact analytical solution for steady, of Newtonian fluids with constant properties. The derivation begins with the simplified Navier-Stokes equations under the assumptions of (no azimuthal or radial velocity components), axisymmetric conditions, fully developed flow (negligible entrance effects), a circular cross-section, and no-slip at the wall. Balancing the axial \frac{dp}{dx} with the viscous shear stress gradient yields \frac{dp}{dx} = \mu \frac{1}{r} \frac{d}{dr} \left( r \frac{du}{dr} \right), where u(r) is the axial . Integrating twice with boundary conditions u(R) = 0 (no-slip) and \frac{du}{dr}(0) = 0 (symmetry) results in the parabolic velocity profile u(r) = \frac{\Delta P}{4 \mu L} (R^2 - r^2), where \Delta P is the pressure drop over length L, and R = D/2 is the tube radius. The average velocity is V = u_{\max}/2 = \frac{\Delta P R^2}{8 \mu L}, and the wall shear stress is \tau_w = -\frac{\Delta P R}{2 L}. The Fanning friction factor is defined as f = \frac{\tau_w}{(\rho V^2)/2}. Substituting the expressions for \tau_w and V gives \Delta P / L = \frac{32 \mu V}{D^2}, which, when equated to the friction factor form \Delta P / L = 2 f \frac{\rho V^2}{D}, yields f = \frac{16}{\mathrm{[Re](/page/Re)}}. This formula applies only for Reynolds numbers \mathrm{[Re](/page/Re)} < 2000 to $2300, beyond which the flow transitions to turbulent. Additionally, the flow must be fully developed, which requires a sufficient entrance length L_e \approx 0.06 \mathrm{[Re](/page/Re)} \, D to establish the parabolic profile.

Non-Circular Ducts

For non-circular ducts in laminar flow, the Fanning friction factor is typically approximated using the hydraulic diameter concept to extend the circular tube baseline, where the exact relation is f = 16 / \mathrm{Re}. The hydraulic diameter D_H is defined as D_H = 4A / P, with A as the cross-sectional area and P as the wetted perimeter; the modified Reynolds number is then \mathrm{Re}_H = \rho u_m D_H / \mu, where u_m is the mean velocity, \rho is the fluid density, and \mu is the dynamic viscosity. The approximation yields f \approx 16 / \mathrm{Re}_H, which provides reasonable estimates for pressure drop calculations in engineering applications. This approach works well for many regular shapes but requires exact solutions or corrections for precise values. For square ducts, the exact Poiseuille solution gives f = 14.227 / \mathrm{Re}_H, derived from series expansions or point-matching methods, representing a deviation of about 11% from the circular approximation. Annular ducts, formed by concentric cylinders with inner-to-outer radius ratio r^*/r_o, exhibit f \mathrm{Re}_H values ranging from 16 (approaching circular as r^*/r_o \to 0) to 24 (as r^*/r_o \to 1, resembling parallel plates), necessitating shape-specific factors from linearization or numerical solutions for accuracy. Representative examples illustrate further adaptations. For infinite parallel plates separated by distance $2h (yielding D_H = 4h), the exact relation is f = 24 / \mathrm{Re}_H, obtained analytically from the velocity profile solution. Equilateral triangular ducts require numerical corrections, with f \mathrm{Re}_H = 13.333 from least-squares fitting, about 17% below the circular value, highlighting the need for shape-dependent adjustments in highly angular geometries. The hydraulic diameter approximation is accurate to within 5% for many common shapes like squares and rectangles but becomes less precise for highly irregular geometries, such as those with sharp corners or varying curvatures, where secondary flows or numerical simulations are recommended for refined predictions.

Formulas for Turbulent Flow

Hydraulically Smooth Pipes

In hydraulically smooth pipes, the wall roughness is negligible relative to the pipe diameter, such that the relative roughness ε/D approaches zero, rendering the Fanning friction factor dependent solely on the Reynolds number Re for fully developed turbulent flow of Newtonian fluids in circular cross-sections. This regime typically applies for Re > 4000, following the transition from which occurs around Re ≈ 2300. The smooth-wall limit assumes no influence from surface protrusions on the turbulent , allowing simplified correlations derived from experimental data and boundary-layer theory. A seminal empirical for the Fanning friction factor in this regime was proposed by Blasius in , given by f = 0.079 \, \mathrm{Re}^{-0.25} valid for $4 \times 10^3 < \mathrm{Re} < 10^5. This power-law relation provides an explicit approximation well-suited to moderate Reynolds numbers, capturing the decrease in friction with increasing flow speed due to reduced relative viscous effects in the turbulent core. For broader applicability across higher Reynolds numbers, the Prandtl-Karman law offers a universal implicit relation for smooth walls, derived from the logarithmic velocity profile in the wall layer: \frac{1}{\sqrt{f}} = 2.0 \log_{10} (\mathrm{Re} \sqrt{f}) - 0.8. This equation, accurate up to Re ≈ 10^7 and beyond, requires iterative solution but aligns closely with experimental measurements by integrating the universal log-law behavior near the wall with the defect law in the outer flow. An alternative explicit approximation extending the range to Re < 10^7 was introduced by Koo in 1933: f = 0.0014 + 0.125 \, \mathrm{Re}^{-0.32}. This formula incorporates a constant term to better fit data at higher Re, where the pure power-law deviates slightly, while maintaining simplicity for engineering calculations in fully developed smooth-pipe flow.

Rough Pipes and General Correlations

In rough pipes, the Fanning friction factor for turbulent flow depends on both the (Re) and the relative roughness ε/D, where ε is the absolute roughness of the pipe inner surface and D is the pipe diameter. Typical values of absolute roughness include approximately 0.046 mm for commercial steel pipes and 0.0015 mm for drawn tubing, reflecting surface irregularities that influence flow resistance. These values lead to relative roughness ε/D that quantifies the roughness effect relative to pipe size, becoming significant in the transition and fully rough regimes of turbulent flow. The Colebrook-White equation provides the standard implicit correlation for the Fanning friction factor in the transition regime between smooth and rough conditions: \frac{1}{\sqrt{f}} = -4 \log_{10} \left( \frac{\epsilon / D}{3.7} + \frac{1.255}{\mathrm{Re} \sqrt{f}} \right) This equation, developed in 1939, requires iterative solution due to the presence of f on both sides and is widely adopted for its accuracy across a broad range of Re and ε/D in turbulent pipe flow. In the fully rough regime, which occurs at high Re or large ε/D (> 0.01) where viscous effects are negligible compared to roughness, the Fanning friction factor simplifies to an explicit form independent of Re: f \approx \frac{1}{[4 \log_{10} (3.7 \, D / \epsilon)]^2} This asymptotic relation derives from the Colebrook-White equation by neglecting the Re-dependent term, capturing the dominance of form drag from surface protrusions. For practical computations avoiding iteration, the Haaland explicit approximation offers a close match to the Colebrook-White equation, particularly for 10^4 < Re < 10^8: \frac{1}{\sqrt{f}} \approx -3.6 \log_{10} \left[ \left( \frac{6.9}{\mathrm{Re}} \right)^{1.11} + \left( \frac{\epsilon / D}{3.7} \right)^{1.11} \right] Proposed in 1983, this formula provides a simple, direct solution with minimal error for engineering applications involving rough pipes. As ε/D approaches 0, these correlations reduce to the hydraulically smooth pipe limit, where roughness effects vanish.

Applications

Pressure Drop Calculations

The frictional pressure drop \Delta P in straight circular pipes due to the Fanning friction factor is given by the formula \Delta P = 4 f \frac{L}{D} \frac{\rho u_m^2}{2}, where f is the Fanning friction factor, L is the pipe length, D is the pipe , \rho is the fluid density, and u_m is the mean flow velocity. This equation arises from a momentum balance accounting for wall and is widely used in chemical and for single-phase incompressible flows. For non-circular ducts, the D is replaced by the D_H = 4A/P, where A is the cross-sectional area and P is the wetted perimeter, ensuring applicability to various geometries while maintaining the same form. Equivalently, the frictional head loss h_f is expressed as h_f = 2 f \frac{L}{D} \frac{u_m^2}{g}, where g is the acceleration due to gravity; this form directly relates to energy requirements, such as pump power P = \rho g Q h_f / \eta, with Q as the volumetric flow rate and \eta as the pump efficiency. The head loss formulation facilitates integration into the mechanical energy balance for systems involving elevation changes and other losses. To compute the pressure drop, follow these steps:
  1. Calculate the Reynolds number \operatorname{Re} = \rho u_m D / \mu, where \mu is the dynamic viscosity, to determine the flow regime (laminar if \operatorname{Re} < 2300, turbulent otherwise).
  2. For laminar flow, use f = 16 / \operatorname{Re}; for turbulent flow, obtain f from the Moody diagram, Colebrook equation \frac{1}{\sqrt{f}} = -2 \log_{10} \left( \frac{\epsilon / D}{3.7} + \frac{1.26}{\operatorname{Re} \sqrt{f}} \right) (with \epsilon as absolute roughness), or explicit approximations.
  3. Substitute f, along with fluid properties, dimensions, and velocity, into the pressure drop formula to yield \Delta P.
Minor losses from fittings, valves, or can be incorporated by adding equivalent lengths L_{eq} to the total L, where L_{eq} / D = K / (4 f) and K is the empirical loss coefficient for the component. For example, consider (\rho = [1000](/page/1000) kg/m³, \mu = 0.001 Pa·s) flowing at u_m = 2 m/s through a 10 cm diameter commercial steel pipe (D = 0.1 m, \epsilon \approx 0.045 mm) over L = 100 m. The is \operatorname{Re} = 200{,}000 (turbulent), yielding f \approx 0.0041 from the Colebrook equation. The is then \Delta P \approx 4 \times 0.0041 \times (100 / 0.1) \times ([1000](/page/1000) \times 2^2 / 2) = 32{,}800 Pa or 33 kPa.

Design of Piping Systems

In the design of piping systems, engineers iterate on diameters to achieve economic velocities, typically ranging from 1 to 3 m/s for liquids, balancing with pumping requirements. This process incorporates the Fanning friction factor f to evaluate total head loss, expressed as H = \sum h_f + minor losses, where frictional components h_f depend on f, pipe length, diameter, and . Such iterations ensure systems operate within limits while minimizing overall lifecycle costs, often starting with initial velocity assumptions refined through . For multiphase flows, such as gas-liquid mixtures common in process industries, the Fanning friction factor is adjusted using the Lockhart-Martinelli parameter X, which quantifies the ratio of liquid-only to gas-only pressure gradients. This parameter enables two-phase multipliers \phi_L^2 and \phi_G^2 to modify single-phase frictional losses calculated with f, improving accuracy for annular or stratified regimes. In high-viscosity oil-gas pipelines, specialized correlations for f in laminar multiphase conditions, derived from mixture Reynolds numbers, enhance pressure drop predictions and system sizing. Integration of the Fanning friction factor into software tools facilitates , particularly for solving implicit relations like the Colebrook equation to obtain f. In (CFD) platforms such as , f is computed from post-simulation, aiding validation of turbulent flow profiles in complex geometries. Spreadsheet solvers in tools like Excel automate Colebrook iterations via goal-seek functions, allowing rapid assessment of f variations with roughness and for preliminary system layouts. In modern applications, the Fanning friction factor informs design for and gas , where it supports modeling to optimize diameters and reduce . For (HVAC) systems, it contributes to duct sizing by relating frictional losses to airflow rates, aligning with standards that limit drops. guidelines incorporate f-based calculations to ensure compliant designs, promoting reduced power and sustainable operation in commercial buildings.

References

  1. [1]
    Concept of Friction Factor in a pipe flow - NPTEL Archive
    The coefficient Cf defined by Eqs (35.1) or (35.3) is known as Fanning's friction factor . To do away with the factor 1/4 in the Eq. (35.3), Darcy defined a ...<|control11|><|separator|>
  2. [2]
    [PDF] Mechanical Energy Balance - Intro and Overview
    Jun 9, 2005 · The Fanning friction factor is defined as a dimensionless wall force in straight tubes, and its relationship to pressure drops, flow rates, and ...
  3. [3]
    6.3.1: Equations Governing Flow in Pipe and Tubing | PNG 301
    The Fanning friction factor is ¼ the value of the Darcy-Weisbach friction factor ... We say that Equation 6.07 is an explicit formula because the friction factor ...
  4. [4]
    FRICTION FACTORS FOR SINGLE PHASE FLOW IN SMOOTH AND ...
    Fanning Friction Factor ... The friction factor is found to be a function of the Reynolds number and the relative roughness. Experimental results of Nikuradse ( ...
  5. [5]
    [PDF] A Glossary of Terms for Fluid Mechanics - University of Notre Dame
    Fanning Friction Factor. Symbol: ff. The wall shear stress divided by the dynamic pressure in turbulent pipe flow. Used for calculating pressure drop in pipes ...
  6. [6]
    [PDF] Pipe Flows - Purdue Engineering
    Dec 15, 2021 · This is the friction factor for turbulent flow in a very rough pipe. The term /D is known as the relative roughness. Note that this equation is ...Missing: Fanning | Show results with:Fanning<|control11|><|separator|>
  7. [7]
  8. [8]
    (PDF) The History of the Darcy-Weisbach Equation for Pipe Flow ...
    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 ...
  9. [9]
    Friction Factor - an overview | ScienceDirect Topics
    The friction factor f is the ratio of the shear stress τ on the bounding walls of the channel to the dynamic pressure of the coolant.
  10. [10]
    [PDF] A Tutorial on Pipe Flow Equations - eng . lbl . gov
    Aug 16, 2001 · The first complication that arises is that there are two common friction factor definitions in standard usage : the Fanning and the Darcy- ...
  11. [11]
    A new explicit friction factor formula for laminar, transition and ...
    John Thomas Fanning (1837–1911) reviewed previous works including Darcy's experiments on friction losses and published tables of f values as a function of pipe ...
  12. [12]
  13. [13]
  14. [14]
    Fanning Friction Factor - an overview | ScienceDirect Topics
    The Fanning friction factor is defined as one fourth the value of the Darcy friction factor and is used in calculations related to pressure drop in fluid flow, ...
  15. [15]
  16. [16]
  17. [17]
    Moody's Friction Factor - an overview | ScienceDirect Topics
    The Darcy friction factor, similar to the Fanning friction factor, can be evaluated by the use of various empirical or theoretical correlations for ...<|control11|><|separator|>
  18. [18]
  19. [19]
    Derivation of Hagen-Poiseuille equation for pipe flows with friction
    Apr 4, 2020 · The Hagen-Poiseuille equation describes the parabolic velocity profile of frictional, laminar pipe flows of incompressible, Newtonian fluids.Missing: Fanning | Show results with:Fanning
  20. [20]
    ME 354 - Lab 6 - Flow in Circular Pipe
    To calculate the flow rate from the velocity profile and from an orifice meter. To calculate the wall shear stress, Darcy friction factor, and average velocity.
  21. [21]
    [PDF] Laminar Flow Forced Convection Heat Transfer and Flow Friction in ...
    Jan 21, 1972 · for an equilateral triangular duct. 145. L. Page 177. Table 27. Equilateral triangular duct with no, one, two and three rounded corners ...
  22. [22]
    [PDF] Fluid mechanics (wb1225) - TU Delft OpenCourseWare
    hydraulic diameter: D h = 4 surface area wetted perimeter flow between 2 parallel plates at distance 2h: f = 24. Re h. ⇒ f = 96. Re. Dh. D h = 4. 2hw. 4h + 2w.
  23. [23]
    [PDF] Friction Factors for Pipe Flow
    The charts apply only to new and clean piping, since the rapid- ity of deterioration with age, dependent upon the quality of the water or fluid and that of the ...Missing: Fanning | Show results with:Fanning
  24. [24]
    Scaling laws for fully developed turbulent flow in pipes - NIH
    The kinematic viscosity of air in the last run which corresponds to a smooth pipe according to the data in ref. 11 can be estimated as 10−2 cm2/s—equal to the ...<|control11|><|separator|>
  25. [25]
    United Formula for the Friction Factor in the Turbulent Region ... - NIH
    May 2, 2016 · 3. Blasius, H. (1913). The law of similarity of frictional processes in fluids. originally in German), ForschArbeitIngenieur-Wesen, Berlin ...
  26. [26]
    A new explicit equation for accurate friction factor calculation of ...
    Blasius (1913) was the first to apply the similarity theory and formulate that the friction factor is a function of the Reynolds number in turbulent flow.
  27. [27]
    "Friction Factors for Pipe Flow" (Moody, Lewis F., 1944, Trans. ASME ...
    14 "The Friction Factors for Clean Round Pipes," by T. B. Drew,. E. C. Koo, and W. H. McAdams, Trans. American Institute of. Chemical Engineers, vol. 28, 1932 ...
  28. [28]
    Absolute Roughness - an overview | ScienceDirect Topics
    Relative Roughness of Pipe ; Commercial steel, 0.0018 ; Wrought iron, 0.0018 ; Asphalted cast iron, 0.0048 ; Galvanized iron, 0.006.
  29. [29]
    Surface Roughness Coefficients - The Engineering ToolBox
    Commercial steel or wrought iron, 0.045 - 0.09, (1.48 - 2.95) 10-4 ... Relative Roughness. Relative roughness - the ratio between absolute roughness an pipe ...
  30. [30]
    Fanning Equation - an overview | ScienceDirect Topics
    The Fanning equation is defined as a method for calculating pressure drop in fluid flow, utilizing the Fanning friction factor, which is one fourth the ...
  31. [31]
    None
    ### Formulas for Pressure Drop and Head Loss Using Fanning Friction Factor
  32. [32]
    [PDF] Rules of Thumb for Chemical Engineers
    For liquids, use 1.5 to 4.0 m/s (5 to 12 ft/s). For dry gas, 15 to 40 m/s (50 to 120 ft/s). For average water service, 1.5 to 3 m/s (5 to 10 ft/s).
  33. [33]
    [PDF] PROCESS DESIGN OF PIPING SYSTEMS (PROCESS PIPING AND ...
    Friction factor of pipe, (dimenssionless) f. D. Darcy's friction factor = f m ... The maximum velocity in piping handling compressible shall be less than ½ of the.
  34. [34]
    [PDF] COMPARISON OF TWO-PHASE PRESSURE DROP MODELS FOR ...
    The most common friction factor is the Fanning friction factor, which is ... where the Lockhart-Martinelli parameter X is defined as: (. ) (. )G. L zp zp. X dd.
  35. [35]
    Frictional Factor Correlation for Laminar High-Viscosity Oil/Gas Flow ...
    Aug 13, 2020 · This paper presents a simplified correlation for two‐phase high‐viscosity oil/gas flow friction factor in horizontal pipes. The developed ...
  36. [36]
    Numerical Results — Lesson 7 | ANSYS Innovation Courses
    The Fanning friction factor, also called the skin friction coefficient, is obtained by non-dimensionalizing the wall shear. It can be calculated and plotted ...
  37. [37]
    [PDF] Pipe Flow/Friction Factor Calculations using Excel Spreadsheets
    Pipe flow calculations use the Darcy-Weisbach and Fanning equations, and the friction factor. These calculations can be done using spreadsheets.
  38. [38]
    [PDF] 2009 ASHRAE Handbook—Fundamentals
    ... Darcy-. Weisbach friction factor. Sometimes a numerically different relation is used with the Fanning friction factor (1/4 of the Darcy friction factor f ).
  39. [39]
    [PDF] ASHRAEs-Duct-System-Design-Guide.pdf - HVAC Simplified
    The variable f is the friction factor and the variable ε is the absolute duct rough- ... The latest editions of energy standards such as ANSI/ASHRAE/IES Standard ...