Friction loss

Last updated

The term friction loss (or frictional loss) has a number of different meanings, depending on its context.

Contents

Jean Le Rond d'Alembert, Nouvelles experiences sur la resistance des fluides, 1777 Alembert - Nouvelles experiences sur la resistance des fluides, 1777 - 14723.jpg
Jean Le Rond d'Alembert, Nouvelles expériences sur la résistance des fluides, 1777

Engineering

Friction loss is a significant engineering concern wherever fluids are made to flow, whether entirely enclosed in a pipe or duct, or with a surface open to the air.

Calculating volumetric flow

In the following discussion, we define volumetric flow rate V̇ (i.e. volume of fluid flowing per time) as

where

r = radius of the pipe (for a pipe of circular section, the internal radius of the pipe).
v = mean velocity of fluid flowing through the pipe.
A = cross sectional area of the pipe.

In long pipes, the loss in pressure (assuming the pipe is level) is proportional to the length of pipe involved. Friction loss is then the change in pressure Δp per unit length of pipe L

When the pressure is expressed in terms of the equivalent height of a column of that fluid, as is common with water, the friction loss is expressed as S, the "head loss" per length of pipe, a dimensionless quantity also known as the hydraulic slope.

where

ρ = density of the fluid, (SI kg / m3)
g = the local acceleration due to gravity;

Characterizing friction loss

Friction loss, which is due to the shear stress between the pipe surface and the fluid flowing within, depends on the conditions of flow and the physical properties of the system. These conditions can be encapsulated into a dimensionless number Re, known as the Reynolds number

where V is the mean fluid velocity and D the diameter of the (cylindrical) pipe. In this expression, the properties of the fluid itself are reduced to the kinematic viscosity ν

where

μ = viscosity of the fluid (SI kg / m  s)

Friction loss in straight pipe

The friction loss in uniform, straight sections of pipe, known as "major loss", is caused by the effects of viscosity, the movement of fluid molecules against each other or against the (possibly rough) wall of the pipe. Here, it is greatly affected by whether the flow is laminar (Re < 2000) or turbulent (Re > 4000): [1]

Form friction

Factors other than straight pipe flow induce friction loss; these are known as "minor loss":

For the purposes of calculating the total friction loss of a system, the sources of form friction are sometimes reduced to an equivalent length of pipe.

Surface roughness

The roughness of the surface of the pipe or duct affects the fluid flow in the regime of turbulent flow. Usually denoted by ε, values used for calculations of water flow, for some representative materials are: [3] [4] [5]

Surface Roughness ε (for water pipes)
Materialmmin
Corrugated plastic pipes (apparent roughness)3.50.14 [6]
Mature foul sewers3.00.12 [6]
Steel water mains with general tuberculations1.20.047 [6]
Riveted Steel0.9–9.00.035–0.35
Concrete (heavy brush asphalts or eroded by sharp material),
Brick
0.50.02 [6] [7]
Concrete0.3–3.00.012–0.12
Wood Stave0.2–0.95–23
Galvanized metals (normal finish),
Cast iron (coated and uncoated)
0.15–0.260.006–0.010 [6]
Asphalted Cast Iron0.120.0048
Concrete (new, or fairly new, smooth)0.10.004 [6]
Steel Pipes, Galvanized metals (smooth finish),
Concrete (new, unusually smooth, with smooth joints),
Asbestos cement,
Flexible straight rubber pipe (with smooth bore)
0.025–0.0450.001–0.0018 [6]
Commercial or Welded Steel, Wrought Iron0.0450.0018
PVC, Brass, Copper, Glass, other drawn tubing0.0015–0.00250.00006–0.0001 [6] [7]

Values used in calculating friction loss in ducts (for, e.g., air) are: [8]

Surface Roughness ε (for air ducts)
Materialmmin
Flexible Duct (wires exposed)3.000.120
Flexible Duct (wires covered)0.900.036
Galvanized Steel0.150.006
PVC, Stainless Steel, Aluminum, Black Iron0.050.0018

Calculating friction loss

Hagen–Poiseuille

Laminar flow is encountered in practice with very viscous fluids, such as motor oil, flowing through small-diameter tubes, at low velocity. Friction loss under conditions of laminar flow follow the Hagen–Poiseuille equation, which is an exact solution to the Navier-Stokes equations. For a circular pipe with a fluid of density ρ and viscosity μ, the hydraulic slope S can be expressed

In laminar flow (that is, with Re < ~2000), the hydraulic slope is proportional to the flow velocity.

Darcy–Weisbach

In many practical engineering applications, the fluid flow is more rapid, therefore turbulent rather than laminar. Under turbulent flow, the friction loss is found to be roughly proportional to the square of the flow velocity and inversely proportional to the pipe diameter, that is, the friction loss follows the phenomenological Darcy–Weisbach equation in which the hydraulic slopeS can be expressed [9]

where we have introduced the Darcy friction factor fD (but see Confusion with the Fanning friction factor);

fD = Darcy friction factor

Note that the value of this dimensionless factor depends on the pipe diameter D and the roughness of the pipe surface ε. Furthermore, it varies as well with the flow velocity V and on the physical properties of the fluid (usually cast together into the Reynolds number Re). Thus, the friction loss is not precisely proportional to the flow velocity squared, nor to the inverse of the pipe diameter: the friction factor takes account of the remaining dependency on these parameters.

From experimental measurements, the general features of the variation of fD are, for fixed relative roughness ε / D and for Reynolds number Re = VD / ν > ~2000, [lower-alpha 1]

The experimentally measured values of fD are fit to reasonable accuracy by the (recursive) Colebrook–White equation, [12] depicted graphically in the Moody chart which plots friction factor fD versus Reynolds number Re for selected values of relative roughness ε / D.

Calculating friction loss for water in a pipe

Water friction loss ("hydraulic slope") S versus flow Q for given ANSI Sch. 40 NPT PVC pipe, roughness height e = 1.5 mm Flow at Constant Friction Loss in PVC pipe.svg
Water friction loss ("hydraulic slope") S versus flow Q for given ANSI Sch. 40 NPT PVC pipe, roughness height ε = 1.5 μm

In a design problem, one may select pipe for a particular hydraulic slope S based on the candidate pipe's diameter D and its roughness ε. With these quantities as inputs, the friction factor fD can be expressed in closed form in the Colebrook–White equation or other fitting function, and the flow volume Q and flow velocity V can be calculated therefrom.

In the case of water (ρ = 1 g/cc, μ = 1 g/m/s [13] ) flowing through a 12-inch (300 mm) Schedule-40 PVC pipe (ε = 0.0015 mm, D = 11.938 in.), a hydraulic slope S = 0.01 (1%) is reached at a flow rate Q = 157 lps (liters per second), or at a velocity V = 2.17 m/s (meters per second). The following table gives Reynolds number Re, Darcy friction factor fD, flow rate Q, and velocity V such that hydraulic slope S = hf / L = 0.01, for a variety of nominal pipe (NPS) sizes.

Volumetric Flow Q where Hydraulic Slope S is 0.01, for selected Nominal Pipe Sizes (NPS) in PVC [14] [15]
NPSDSRefDQV
inmmin [16] gpmlpsft/sm/s
1/2150.6220.0144675.080.90.0550.9280.283
3/4200.8240.0173015.4520.1201.1440.349
1251.0490.01110905.763.80.2321.3660.416
1+1/2401.6100.01231216.32120.7431.8550.565
2502.0670.01353606.64241.4582.2100.674
3753.0680.01688687.15704.2152.8990.884
41004.0260.011086157.501448.7233.4851.062
61506.0650.012150018.0343026.0134.5791.396
82007.9810.013388628.3989253.9515.4841.672
1025010.0200.014933578.68163198.6176.3601.938
1230011.9380.016582548.902592156.7657.1222.171

Note that the cited sources recommend that flow velocity be kept below 5 feet / second (~1.5 m/s).

Also note that the given fD in this table is actually a quantity adopted by the NFPA and the industry, known as C, which has the imperial units psi/(100 gpm2ft) and can be calculated using the following relation:

where is the pressure in psi, is the flow in 100gpm and is the length of the pipe in 100ft

Calculating friction loss for air in a duct

A graphical depiction of the relationship between Dp / L, the pressure loss per unit length of pipe, versus flow volume Q, for a range of choices for pipe diameter D, for air at standard temperature and pressure. Units are SI. Lines of constant Re[?]fD are also shown. Equal-friction chart for air in metal duct (e = 0.05mm).svg
A graphical depiction of the relationship between Δp / L, the pressure loss per unit length of pipe, versus flow volume Q, for a range of choices for pipe diameter D, for air at standard temperature and pressure. Units are SI. Lines of constant RefD are also shown.

Friction loss takes place as a gas, say air, flows through duct work. [17] The difference in the character of the flow from the case of water in a pipe stems from the differing Reynolds number Re and the roughness of the duct.

The friction loss is customarily given as pressure loss for a given duct length, Δp / L, in units of (US) inches of water for 100 feet or (SI) kg / m2 / s2.

For specific choices of duct material, and assuming air at standard temperature and pressure (STP), standard charts can be used to calculate the expected friction loss. [8] [18] The chart exhibited in this section can be used to graphically determine the required diameter of duct to be installed in an application where the volume of flow is determined and where the goal is to keep the pressure loss per unit length of duct S below some target value in all portions of the system under study. First, select the desired pressure loss Δp / L, say 1 kg / m2 / s2 (0.12 in H2O per 100 ft) on the vertical axis (ordinate). Next scan horizontally to the needed flow volume Q, say 1 m3 / s (2000 cfm): the choice of duct with diameter D = 0.5 m (20 in.) will result in a pressure loss rate Δp / L less than the target value. Note in passing that selecting a duct with diameter D = 0.6 m (24 in.) will result in a loss Δp / L of 0.02 kg / m2 / s2 (0.02 in H2O per 100 ft), illustrating the great gains in blower efficiency to be achieved by using modestly larger ducts.

The following table gives flow rate Q such that friction loss per unit length Δp / L (SI kg / m2 / s2) is 0.082, 0.245, and 0.816, respectively, for a variety of nominal duct sizes. The three values chosen for friction loss correspond to, in US units inch water column per 100 feet, 0.01, .03, and 0.1. Note that, in approximation, for a given value of flow volume, a step up in duct size (say from 100mm to 120mm) will reduce the friction loss by a factor of 3.

Volumetric Flow Q of air at STP where friction loss per unit length Δp / L (SI kg / m2 / s2) is, resp., 0.082, 0.245, and 0.816., for selected Nominal Duct Sizes [19] in smooth duct (ε = 50μm.)
Δp / L0.082 0.245 0.816
kg / m2 / s2
Duct size QQQ
inmmcfmm3/scfmm3/scfmm3/s
2+1/26330.001250.0024100.0048
3+1/48050.0024100.0046200.0093
4100100.0045180.0085360.0171
5125180.0083330.0157660.0313
6160350.0163650.03081290.0611
8200640.03011190.05632360.1114
102501170.05512180.10304300.2030
123152180.10314070.19197990.3771
164004160.19657720.364615130.7141
205007590.358214040.662727431.2945
2463014110.665726031.228550722.3939
3280026731.261349192.321795634.5131
40100048472.287789034.2018172708.1504
48120078763.7172144426.81612796913.2000

Note that, for the chart and table presented here, flow is in the turbulent, smooth pipe domain, with R* < 5 in all cases.

Notes

Further reading

Related Research Articles

<span class="mw-page-title-main">Laminar flow</span> Flow where fluid particles follow smooth paths in layers

Laminar flow is the property of fluid particles in fluid dynamics to follow smooth paths in layers, with each layer moving smoothly past the adjacent layers with little or no mixing. At low velocities, the fluid tends to flow without lateral mixing, and adjacent layers slide past one another smoothly. There are no cross-currents perpendicular to the direction of flow, nor eddies or swirls of fluids. In laminar flow, the motion of the particles of the fluid is very orderly with particles close to a solid surface moving in straight lines parallel to that surface. Laminar flow is a flow regime characterized by high momentum diffusion and low momentum convection.

In fluid dynamics, turbulence or turbulent flow is fluid motion characterized by chaotic changes in pressure and flow velocity. It is in contrast to a laminar flow, which occurs when a fluid flows in parallel layers, with no disruption between those layers.

In fluid dynamics, the Darcy–Weisbach equation is an empirical equation that relates the head loss, or pressure loss, due to friction along a given length of pipe to the average velocity of the fluid flow for an incompressible fluid. The equation is named after Henry Darcy and Julius Weisbach. Currently, there is no formula more accurate or universally applicable than the Darcy-Weisbach supplemented by the Moody diagram or Colebrook equation.

<span class="mw-page-title-main">Parasitic drag</span> Aerodynamic resistance against the motion of an object

Parasitic drag, also known as profile drag, is a type of aerodynamic drag that acts on any object when the object is moving through a fluid. Parasitic drag is a combination of form drag and skin friction drag. It affects all objects regardless of whether they are capable of generating lift.

<span class="mw-page-title-main">Bingham plastic</span> Material which is solid at low stress but becomes viscous at high stress

In materials science, a Bingham plastic is a viscoplastic material that behaves as a rigid body at low stresses but flows as a viscous fluid at high stress. It is named after Eugene C. Bingham who proposed its mathematical form.

An orifice plate is a device used for measuring flow rate, for reducing pressure or for restricting flow.

<span class="mw-page-title-main">Plug flow</span> Simple model of fluid flow in a pipe

In fluid mechanics, plug flow is a simple model of the velocity profile of a fluid flowing in a pipe. In plug flow, the velocity of the fluid is assumed to be constant across any cross-section of the pipe perpendicular to the axis of the pipe. The plug flow model assumes there is no boundary layer adjacent to the inner wall of the pipe.

The Fanning friction factor, named after John Thomas Fanning, is a dimensionless number used as a local parameter in continuum mechanics calculations. It is defined as the ratio between the local shear stress and the local flow kinetic energy density:

The Hazen–Williams equation is an empirical relationship which relates the flow of water in a pipe with the physical properties of the pipe and the pressure drop caused by friction. It is used in the design of water pipe systems such as fire sprinkler systems, water supply networks, and irrigation systems. It is named after Allen Hazen and Gardner Stewart Williams.

<span class="mw-page-title-main">Law of the wall</span> Relation of flow speed to wall distance

In fluid dynamics, the law of the wall states that the average velocity of a turbulent flow at a certain point is proportional to the logarithm of the distance from that point to the "wall", or the boundary of the fluid region. This law of the wall was first published in 1930 by Hungarian-American mathematician, aerospace engineer, and physicist Theodore von Kármán. It is only technically applicable to parts of the flow that are close to the wall, though it is a good approximation for the entire velocity profile of natural streams.

The Ergun equation, derived by the Turkish chemical engineer Sabri Ergun in 1952, expresses the friction factor in a packed column as a function of the modified Reynolds number.

In respiratory physiology, airway resistance is the resistance of the respiratory tract to airflow during inhalation and exhalation. Airway resistance can be measured using plethysmography.

Johann Nikuradse was a Georgia-born German engineer and physicist. His brother, Alexander Nikuradse, was also a Germany-based physicist and geopolitician known for his ties with Alfred Rosenberg and for his role in saving many Georgians during World War II.

The Dean number (De) is a dimensionless group in fluid mechanics, which occurs in the study of flow in curved pipes and channels. It is named after the British scientist W. R. Dean, who was the first to provide a theoretical solution of the fluid motion through curved pipes for laminar flow by using a perturbation procedure from a Poiseuille flow in a straight pipe to a flow in a pipe with very small curvature.

In fluid dynamics, the Darcy friction factor formulae are equations that allow the calculation of the Darcy friction factor, a dimensionless quantity used in the Darcy–Weisbach equation, for the description of friction losses in pipe flow as well as open-channel flow.

In fluid mechanics, pipe flow is a type of fluid flow within a closed conduit, such as a pipe, duct or tube. It is also called as Internal flow. The other type of flow within a conduit is open channel flow. These two types of flow are similar in many ways, but differ in one important aspect. Pipe flow does not have a free surface which is found in open-channel flow. Pipe flow, being confined within closed conduit, does not exert direct atmospheric pressure, but does exert hydraulic pressure on the conduit.

<span class="mw-page-title-main">Moody chart</span> Graph used in fluid dynamics

In engineering, the Moody chart or Moody diagram is a graph in non-dimensional form that relates the Darcy–Weisbach friction factor fD, Reynolds number Re, and surface roughness for fully developed flow in a circular pipe. It can be used to predict pressure drop or flow rate down such a pipe.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Hagen in 1839 and then by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

<span class="mw-page-title-main">Reynolds number</span> Ratio of inertial to viscous forces acting on a liquid

In fluid dynamics, the Reynolds number is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be dominated by laminar (sheet-like) flow, while at high Reynolds numbers, flows tend to be turbulent. The turbulence results from differences in the fluid's speed and direction, which may sometimes intersect or even move counter to the overall direction of the flow. These eddy currents begin to churn the flow, using up energy in the process, which for liquids increases the chances of cavitation.

In fluid dynamics, the entrance length is the distance a flow travels after entering a pipe before the flow becomes fully developed. Entrance length refers to the length of the entry region, the area following the pipe entrance where effects originating from the interior wall of the pipe propagate into the flow as an expanding boundary layer. When the boundary layer expands to fill the entire pipe, the developing flow becomes a fully developed flow, where flow characteristics no longer change with increased distance along the pipe. Many different entrance lengths exist to describe a variety of flow conditions. Hydrodynamic entrance length describes the formation of a velocity profile caused by viscous forces propagating from the pipe wall. Thermal entrance length describes the formation of a temperature profile. Awareness of entrance length may be necessary for the effective placement of instrumentation, such as fluid flow meters.

References

  1. 1 2 Munson, B.R. (2006). Fundamentals of Fluid Mechanics (5 ed.). Hoboken, NJ: Wiley & Sons.
  2. Allen, J.J.; Shockling, M.; Kunkel, G.; Smits, A.J. (2007). "Turbulent flow in smooth and rough pipes". Phil. Trans. R. Soc. A. 365 (1852): 699–714. Bibcode:2007RSPTA.365..699A. doi:10.1098/rsta.2006.1939. PMID   17244585. S2CID   2636599. Per EuRoPol GAZ website.
  3. "Pipe Roughness". Pipe Flow Software. Retrieved 5 October 2015.
  4. "Pipe Roughness Data". Efunda.com. Retrieved 5 October 2015.
  5. "Pipe Friction Loss Calculations". Pipe Flow Software. Retrieved 5 October 2015. The friction factor C in the Hazen-Williams formula takes on various values depending on the pipe material, in an attempt to account for surface roughness.
  6. 1 2 3 4 5 6 7 8 Chung, Yongmann. "ES2A7 laboratory Exercises" (PDF). University of Warwick, School of Engineering. Retrieved 20 October 2015.
  7. 1 2 Sentürk, Ali. "Pipe Flow" (PDF). T.C. İSTANBUL KÜLTÜR UNIVERSITY. Retrieved 20 October 2015.
  8. 1 2 "On-Line Duct Friction Loss". FreeCalc.com. Retrieved 8 October 2015.
  9. Brown, G.O. (2003). "The History of the Darcy-Weisbach Equation for Pipe Flow Resistance". Environmental and Water Resources History. American Society of Civil Engineers. pp. 34–43. doi:10.1061/40650(2003)4.
  10. Nikuradse, J. (1933). "Strömungsgesetze in Rauen Rohren". V. D. I. Forschungsheft. 361: 1–22.
  11. Moody, L. F. (1944), "Friction factors for pipe flow", Transactions of the ASME, 66 (8): 671–684
  12. Rao, A.; Kumar, B. "Friction Factor for Turbulent Pipe Flow" (PDF). Retrieved 20 October 2015.
  13. "Water - Dynamic and Kinetic Viscosity". Engineering Toolbox. Retrieved 5 October 2015.
  14. "Technical Design Data" (PDF). Orion Fittings. Retrieved 29 September 2015.
  15. "Tech Friction Loss Charts" (PDF). Hunter Industries. Retrieved 5 October 2015.
  16. "Pipe Dimensions" (PDF). Spirax Sarco Inc. Retrieved 29 September 2015.
  17. 1 2 Elder, Keith E. "Duct Design" (PDF). Retrieved 8 October 2015.
  18. Beckfeld, Gary D. (2012). "HVAC Calculations and Duct Sizing" (PDF). PDH Online, 5272 Meadow Estates Drive Fairfax, VA 22030. Retrieved 8 October 2015.
  19. "Circular Duct Sizes". The Engineering Toolbox. Retrieved 25 November 2015.