Friction Factor

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 79

Learn more about Friction Factor

A Review of Production Engineering Fundamentals


Gabor Takacs PhD, in Sucker-Rod Pumping Handbook, 2015

2.3.2.5 Moody diagram

Friction factor, in general, was found to be a function of the Reynolds number and pipe relative
roughness. Relative roughness is defined here as the ratio of the absolute roughness ε of the pipe
inside wall to the pipe inside diameter:

(2.34)k=εd

where:

k = pipe relative roughness, –,

ε = pipe absolute roughness, in,

d = pipe diameter, in.

Typical absolute roughness values are ε = 0.0006 in for new and ε = 0.009 in for used well
tubing.

There are several formulae describing friction factors for different flow conditions. Most of them
are included in the Moody diagram [18], which is a graphical presentation of Darcy–Weisbach-
type f values. Use of this chart, given in Fig. 2.6, is generally accepted in the petroleum industry.
Sign in to download full-size image

Figure 2.6. The Moody [18] diagram: friction factors for pipe flow.

As seen in the Moody diagram, the friction factor is a different function of the variables NRe and
k in different ranges. In laminar flow (i.e., Reynolds numbers lower than 2,000–2,300) f varies
with the Reynolds number only:

(2.35)f=64NRe

On the other hand, in fully developed turbulent flow, for Reynolds numbers over 2,000–2,300
and for a rough pipe, friction factor is a sole function of the pipe's relative roughness. There
exists a transition region between smooth wall flow and fully developed turbulent flow where the
value of friction factor is determined by both NRe and k values.

For computer applications the use of the Colebrook equation is recommended; it can be solved
for friction factors in both the transition and the fully developed turbulent regions [19]. It is an
implicit function and can be solved for f values using some iterative procedure, e.g., the
Newton–Raphson scheme.

(2.36)1f=2log3.7k−2log(1+9.335kNRef)

where:

f = friction factor, –,

k = pipe relative roughness, –,

NRe = Reynolds number, –.


The inconvenience of an iteration scheme can be eliminated if an explicit formula is used to
calculate friction factors. Gregory and Fogarasi [20] investigated several such models and
found that the formula developed by Chen [21] gives very good results. This formula, as given
below, very accurately reproduces the Moody diagram over the entire range of conditions. It
does not necessitate an iterative scheme and thus can speed up lengthy calculations (e.g.,
multiphase pressure drop calculations) involving the determination of a great number of friction
factors.

(2.37)1f=−2log(k3.7065−5.0452NRelogA)

where:

(2.38)A=k1.10982.8257+(7.149NRe)0.8981

Hydrodynamic Stability of Pipelines


Boyun Guo, ... Tian Ran Lin, in Offshore Pipelines (Second Edition), 2014

4.3.1.4 Soil Friction Factor

Friction factor (μ) is defined as the ratio between the force required to move a section of pipe and
the vertical contact force applied by the pipe on the seabed. This simplified model (Coulomb) is
used to assess stability. The friction factor depends on the type of soil, the pipe roughness,
seabed slope, and depth of burial. For practical purposes, only the type of soil is considered and
the pipe roughness ignored.

For stability analysis, a lower bound estimate for soil friction is conservatively assumed, whereas
for pulling or towing analysis, an upper bound estimate would be appropriate. The following
lateral friction factors are given as guidelines for stability analysis in the absence of site-specific
data:

Loose sand: μ=tan ϕ (generally ϕ=30°)

Compact sand: μ=tan ϕ (generally ϕ=35°)

Soft clay: μ=0.7

Stiff clay: μ=0.4

Rock and gravel: μ=0.7.

These coefficients represent the “best” estimate for generalized soil types and do not include
safety factors.

Small scale tests (Lyons, 1973) and offshore tests (Lambrakos, 1985) have shown that the
starting friction factor in sand is about 30% less than the maximum value which occurs after a
very small displacement of the pipe builds a wedge of soil; past this point, the friction factor
levels off. The value given above accounts for the build-up of this wedge of soil which has been
shown to take place. The Coulomb model underestimates the actual lateral soil resistance if
settlement is anticipated.

Design and Construction of Offshore Pipelines


André C. Nogueira, David S. Mckeehan, in Handbook of Offshore Engineering, 2005

11.8.1 Soil Friction Factor

The friction factor is defined as the ratio between the force required to move a section of pipe
and the vertical contact force applied by the pipe on the seabed. This simplified model
(Coulomb) is used to assess stability and requires an estimate of the friction factor,. Strictly
speaking, the friction factor,, depends on the type of soil, the pipe roughness, seabed slope and
depth of burial; however, the pipe roughness is typically ignored.

For stability analysis, a lower bound estimate for soil friction is conservatively assumed, whereas
for pulling or towing analysis, an upper bound estimate would be appropriate. The following
lateral friction factors [Lyons, 1973; Lambrakos, 1985] are given as a guideline for stability
analysis in the absence of site-specific data:

•Loose sand: tan Φ (generally Φ = 30°)


•Compact sand: tan Φ (generally Φ = 35°)
•Soft clay: 0.7
•Stiff clay: 0.4
•Rock and gravel: 0.7

These coefficients are adequate for generalised soil types and do not include safety factors.
Small-scale tests [Lyons, 1973] and offshore tests [Lambrakos, 1985] have shown that the
starting friction factor in sand is about 30% less than the maximum value, which occurs after a
very small displacement of the pipe builds a wedge of soil; past this point, the friction factor
levels off. The values given above account for the build-up of this wedge of soil, which has been
shown to take place.

Read full chapter

Fluid Flow
R.H.S. WINTERTON, in Thermal Design of Nuclear Reactors, 1981

The friction factor

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, i.e.
(6.1)τ=f12ρu2

where ρ is the coolant density and u its velocity (the dynamic pressure 1/2ρu2 is the rise in
pressure that ideally could be obtained if the coolant were slowed down to zero velocity).

An alternative friction factor sometimes used is the Darcy—Weisbach friction factor defined by

f′=τ/2ρu2=f/4

The friction factor is dimensionless, and for turbulent flow over a given type of surface it is
roughly constant, being only weakly dependent on Reynolds number and channel geometry.

Foam Drilling
Bill Rehm, ... Amir Paknejad, in Underbalanced Drilling: Limits and Extremes, 2012

4.39 Foam Friction Factor

Determining friction factor is crucial for foam flow calculations. Assuming that stable foam flow
falls into a laminar flow regime, the theoretical approach for the Moody friction factor is
expressed as a function of Reynolds number:

(4.29)f=64Re

with Reynolds number calculated as

(4.30)NRe=v¯DHρ¯μe

where

f = Moody friction factor

NRe = Reynolds number

v¯ = Average foam velocity

ρ¯ = Average foam density

μe = Effective viscosity

It has been reported in several cases that the friction factor given by Eq. (4.29) has been too high.
Guo et al.15 developed an empirical correlation derived from two-phase flow regimes that gives
good results for foam flow in conditions commonly encountered in foam drilling. This empirical
approach, which uses the weight flow rate, is expressed as

(4.31)w=0.0765sSgQg+8.33SlQl
where

w = Mass flow rate

(4.32)Dρv=0.02173wDH
(4.33)f=4×101.444−2.5log(Dρv)

For the Reynolds numbers ranging 2,000–4,000, transitional flow conditions, the Colbrook-
White equation can be used to calculate the friction factor. This equation can not be solved
directly and must be solved by trial and error.

(4.34)1f=−2log(ε3.7⋅D+2.51NRe⋅f)

The Churchill equation is another empirical equation developed for the calculation of the friction
factor in transitional flow conditions. The Churchill equation is represented as

(4.35)f=8⋅((8NRe)12+1(A+B)1.5)112

where

(4.36)A=(2.457⋅Ln(1(7NRe)0.9+0.27⋅εD)) 16
(4.37)B=(37530NRe)16

When the Reynolds number is greater than 4,000, turbulent flow regime, the smooth pipe friction
factor can be calculated using a Blasius type equation as

(4.38)f=0.184(NRe)−0.2

At large Reynolds numbers, the friction factor is independent of fluid viscosity and only depends
on the roughness of the pipe. The Nikuradse equation for fully rough flow can be used to
calculate the friction factor as

(4.39)1f=1.74−2Log(2⋅εD)

Jain came up with a non-iterative method that covers all ranges of smooth, transition, and fully
rough flow. The Jain equation is represented as

(4.40)1f=1.14−2Log(εD⋅21.25NRe0.9)

Pressure Loss through Piping Systems


E. Shashi Menon P.E., Pramila S. Menon MBA, in Working Guide to Pumps and Pumping
Stations, 2010

Determining the Friction Factor from the Moody Diagram


The friction factor f can be determined using the Moody diagram shown in Figure 4.2 as follows:

1.For the given flow rate, liquid properties, and pipe size, calculate the Reynolds number of flow
using Equation (4.8).
2.Calculate the relative roughness (e/D) of the pipe by dividing the pipe absolute roughness by
the inside diameter of the pipe.
3.Starting at the Reynolds number value on the horizontal axis of the Moody diagram, Figure
4.2, move vertically up to the relative roughness curve. Then move horizontally to the left and
read the friction factor f on the vertical axis on the left.
Example 4.8

USCS Units

Using the Moody diagram, determine the friction factor for a crude oil pipeline with a 16-inch
outside diameter and a 0.250-inch wall thickness at a flow rate of 6250 bbl/h. Viscosity of the
crude oil is 15.0 cSt. The absolute pipe roughness = 0.002 in.

Solution

The inside diameter of pipe D = 16 − (2 × 0.250) = 15.5 in. Using Equation 4.9, we calculate the
Reynolds number as follows:

R=(2214×6250)/(15.5×15)=59,516

Relative roughness = 0.002/15.5 = 0.000129. From the Moody diagram, for R = 59,516 and
(e/D) = 0.000129, we get the friction factor as f = 0.0206.

Example 4.9

SI Units

Using the Moody diagram, determine the friction factor for a water pipeline with a 400 mm
outside diameter and a 6 mm wall thickness at a flow rate of 400 m3/h. Viscosity of water is
1.0 cSt. The absolute pipe roughness = 0.05 mm.

Solution

The inside diameter of pipe D = 400 − (2 × 6) = 388 mm. Using Equation 4.10, we calculate the
Reynolds number as follows:

R=(353,678×400)/(1×388)=364,617

Relative roughness = 0.05/388 = 0.000129. From the Moody diagram, for R = 364,617 and (e/D)
= 0.000129, we get the friction factor as f = 0.0153.

Read full chapter


Pipeline Hydraulic Analysis
E. Shashi Menon Ph.D., P.E., in Pipeline Planning and Construction Field Manual, 2011

8.7 Moody Diagram

The friction factor equations discussed in the preceding section are also plotted on the Moody
diagram as shown in Fig. 8.1. This diagram shows the friction factor on the left vertical axis,
plotted against the Reynolds number on the horizontal axis, for various values of the relative
roughness of the pipe (e/D).

Sign in to download full-size image

Figure 8.1. Moody diagram for friction factor.

The Moody diagram represents the complete friction factor map for laminar and all turbulent
regions of pipe flows. It is commonly used in estimating friction factor in pipe flow. If the
Moody diagram is not available, we must use a trial and error solution of Eq. (8.18) to calculate
the friction factor.

To use the Moody diagram for determining the friction factor f we first calculate the relative
roughness (e/D) and the Reynolds number R for the flow. Next, for this value of R on the
horizontal axis, draw a vertical line that intersects with the appropriate relative roughness (e/D)
curve. From this point of intersection on the (e/D) curve, we go horizontally to the left and read
the value of the friction factor f on the vertical axis on the left.
Note that some publications refer to the Fanning friction factor. This is equal to one-fourth the
value of the Darcy friction factor f discussed in this chapter. Unless otherwise specified, we will
use the Darcy friction factor f throughout this book.

Table 8.6 shows the variation of the friction factor for various Reynolds numbers for three
relative roughness values.

Table 8.6. Darcy Friction Factors*

Friction Factor
Reynolds Number
e/D = 0.0001 e/D = 0.0002 e/D = 0.0003
1000 0.0664 0.0756 0.0799
2000 0.0512 0.0678 0.0762
4000 0.0407 0.0618 0.0733
5000 0.0380 0.0602 0.0725
10,000 0.0311 0.0559 0.0703
20,000 0.0261 0.0524 0.0684
30,000 0.0237 0.0506 0.0675
40,000 0.0222 0.0495 0.0669
50,000 0.0212 0.0487 0.0664
60,000 0.0204 0.0481 0.0661
70,000 0.0198 0.0476 0.0658
80,000 0.0192 0.0471 0.0656
90,000 0.0188 0.0468 0.0654
100,000 0.0185 0.0465 0.0652
125,000 0.0177 0.0458 0.0648
150,000 0.0172 0.0454 0.0646
200,000 0.0164 0.0447 0.0642
225,000 0.0161 0.0444 0.0640
250,000 0.0158 0.0442 0.0639
275,000 0.0156 0.0440 0.0638
300,000 0.0154 0.0438 0.0637
325,000 0.0153 0.0436 0.0636
350,000 0.0151 0.0435 0.0635
375,000 0.0150 0.0434 0.0634
400,000 0.0149 0.0433 0.0634
425,000 0.0147 0.0432 0.0633
450,000 0.0146 0.0431 0.0632
500,000 0.0145 0.0429 0.0631
750,000 0.0139 0.0423 0.0628
1,000,000 0.0135 0.0419 0.0626
*
Friction factor based on Swamee–Jain equation: f = 0.25/[Log10(e/3.7D + 5.74/R0.9)]2.
Example Problem 8.2 (USCS)

Water (Sg = 1.0 and visc = 1.0 cSt) flows through an NPS 20, 0.375 in. wall pipe at 6000
gal/min. Calculate the friction factor using the Colebrook–White equation. Assume an absolute
pipe roughness of 0.002 in. What is the head loss due to friction in 3500 ft of pipe?

Solution

First, we calculate the Reynolds number from Eq. (8.8) as follows:

R=3160×6000/(19.25×1.0)=984,935

Since R > 4000, the flow is fully turbulent and the friction factor f is calculated using Eq. (8.18)
as follows:

1/√f=−2 Log10[(0.002/(3.7×19.25))+2.51/(984,935√f)]

The above implicit equation for f must be solved by trial and error.

First assume a trial value of f = 0.02. Substituting in the equation above, we get successive
approximations for f as follows:

f=0.0133, 0.0135, and 0.0135

Therefore, the solution is f = 0.0135.

Converting the given flow rate from gal/min to bbl/day,

Q=6000 gal/min×(60×24 min/day)×(1/42 gal/bbl)=205,714.29 bbl/day

Using Eq. (8.14), the pressure drop resulting from friction is

Pm=0.0605×0.0135×(205,714.29)2(1.0/19.255)=13.08 psi/mi

Therefore, the pressure drop in 3500 ft of pipe is

ΔP=13.08×(3500/5280)=8.67 psi

This is converted to head in feet of water, using Eq. (8.10):

Head loss due to friction=8.67×2.31/1.0=20.03 ft


Example Problem 8.3 (SI)

Diesel fuel (Sg = 0.85 and visc = 5.0 cSt) flows in a pipeline, DN 400, 8 mm wall thickness at
580 m3/h. Calculate the average velocity, Reynolds number, and the friction factor. Assume an
absolute pipe roughness of 0.05 mm. What is the pressure drop resulting from friction in 5 km
length of the pipeline?

Solution

The inside diameter of the pipe, D = 400 − 2 × 8 = 384 mm.

The average flow velocity is found using Eq. (8.3):

V=353.6777(580)/3842=1.39 m/s

Next, we calculate the Reynolds number from Eq. (8.9) as follows:

R=353,678 (580)/(5×384)=106,840

Because R > 4000, the flow is fully turbulent and the friction factor f is calculated using Eq.
(8.19) as follows:

1/√f=−2Log10[(0.05/(3.7×384))+2.51/(106,840√f)]

This equation for f must be solved by trial and error.

First, assume a trial value of f = 0.02. Substituting in the equation above, and by successive
iteration we get f = 0.0612.

The pressure drop due to friction is calculated from Eq. (8.16) as

Pkm=6.2475×1010×0.0612×(580)2(0.85/3845)=130.94 kPa/km
The pressure drop in 5 km length of the pipeline=130.94×5=654.7 kPa

Fluid Flow
A. Kayode Coker, in Fortran Programs for Chemical Process Design, Analysis, and Simulation,
1995

Friction Factor

The friction factor is related to the Reynolds number by a set of correlations and depends on
whether the flow regime is laminar, transitional, or turbulent. The Reynolds number is:

(3-12)NRe=ρVDμ=50.6Qρdμ=5.31Wdμ

where Q = volumetric flowrate, gals/min

W = Mass flow rate, lb/h


μ = fluid viscosity, cP

For laminar flow with NRe ≤ 2000

(3-13)fD=64NRe

The Darcy friction factor is four times the Fanning friction factor, fF, i.e., fD = 4fF. For fully
developed turbulent flow regime in smooth and rough pipes, the Colebrook [5] equation or the
Chen [6] equation can be used.

The Colebrook equation is expressed as:

(3-14)1fD=−0.8686 ln{ɛ3.7D+2.51NRefD}

Equation 3-14 is implicit in fD, as it cannot be rearranged to derive fD directly and thus requires
an iterative solution. Here, the Chen equation is used for calculating fD, (i.e, fD = 4fc). The Chen
equation is explicit and easier to use than Equation 3-14). This can be expressed as:

(3-15)1fC=−4log{ɛ3.7D−5.02NRelogA}

where

A=ɛ/D3.7+(6.7NRe)0.9

and

ɛ = pipe roughness, ft

A detailed review of other explicit equations is given by Gregory and Fogarasi [7]. Different
piping materials are often used in the chemical process industries, and at a high Reynolds
number, the friction factor is affected by the roughness of the surface. This is measured as the
ratio ɛ/D of projections on the surface to the diameter of the pipe. Glass and plastic pipe
essentially have ɛ = 0. Values of ɛ are shown in Table 3-3.

Table 3-3. Values of Absolute Pipe Roughness

Pipe Material ɛ ft
Riveted steel 0.003-0.03
Concrete 0.001-0.01
Wood stave 0.0006-0.003
Cast iron 0.00085
Galvanized iron 0.0005
Asphalted cast iron 0.0005
Commercial steel or wrought iron 0.00015
Drawn tubing 0.000005
Analytical Solutions to Poiseuille Flow Problems in Different
Geometries
Bastian E. Rapp, in Microfluidics: Modelling, Mechanics and Mathematics, 2017

16.3.5 Darcy Friction Factor

Darcy Friction Factor for Laminar Hagen-Poiseuille Flow. You may occasionally come
across a notation of Eq. 16.34 that can be derived by rearrangement and is given as

(Eq. 16.37)dpdz=−vav8ηR2ReRe=−vav32ηd22Re2Re=−vav64ηd2Re2Re=−vavfη2R2Re

where we have introduced a factor that is referred to as the Darcy friction factor f for Hagen-
Poiseuille flow. It is given by

(Eq. 16.38)fHagen-Poiseuille=64Re

The Darcy friction factors are experimentally determined coefficients that allow a correlation
between average flow velocity (and therefore volume flowrate) and pressure drop expressed
solely by constants and the Reynolds number. This you can see from Eq. 16.37. Eq. 16.38 is the
friction factor for laminar Hagen-Poiseuille flow. It is only valid for Reynolds numbers in
laminar flow. If the Reynolds number increases to the transition regime or even to turbulent flow
Eq. 16.37 can still be applied, but a more complicated equation for the Darcy friction factor
would need to be found. As stated, these are usually derived empirically from experimental data.
By using only complicated equations for this one factor the overall equations remain very
simple. Eq. 16.37 is generally referred to as the Darcy-Weisbach1,2equation.

Higher Reynolds Numbers: Colebrook-White Equation. If Reynolds numbers are above 4000
different approximations for the Darcy friction factor have to be found. Most common is the
following implicit formula that was first given by Colebrook1[5] based on earlier work by him
and White [6], which is why it is commonly referred to as the Colebrook-White equation:

(Eq. 16.39)1f=−2log10(ε3.7dH+2.51Ref)

which uses the roughness height ε and the hydraulic diameter (see section 16.3.6). Eq. 16.39 is
an implicit formula, meaning that it needs to be solved numerically as there is no analytical
solution for f. However, this equation can be approximated conveniently using algebra tools or
even numerical calculators. However, several attempts to finding explicit analytical
approximations to this equation for certain fluid mechanical cases have been made, which
resulted in a wide range of approximations for f from Eq. 16.39. This work is still ongoing with
recent contributions [7].

Read full chapter

Pressure Drop
Sujoy K. Saha, Gian P. Celata, in Microchannel Phase Change Transport Phenomena, 2016

5.5.2 Effect of Aspect Ratio on Pressure Drop

Dimensionless parameter, Poiseuille number, which is a product of friction factor and Reynolds
number, has been used as an indicator of pressure drop.

The friction factor is calculated using the Darcy equation [49] and is given by

f=2DhΔpρu2Lc

where Dh is the hydraulic diameter, Δp is the pressure drop, ρ is the density of fluid, u is the
velocity of fluid, and Lc is the length of channel.

The effect of aspect ratio of rectangular microchannel on Poiseuille number is demonstrated in


Fig. 5.20. Pressure drop is observed to increase with increase in aspect ratio. The same
observation was reported by Kandlikar et al. [50] for rectangular channels. The increase in
pressure drop is attributed to vortices that are induced due to reduced free flow passage with
increase in aspect ratio.

Sign in to download full-size image

Figure 5.20. Poiseuille number at different width–height (Wc/Hc) ratios for rectangular
microchannels [47].

Reynolds Number - the non-dimensional velocity - can be defined as the ratio

 inertia force (ρ u L) to viscous or friction force (μ)

and interpreted as the ratio

 dynamic pressure (ρ u2) to shearing stress (μ u / L)


Reynolds Number can therefore be expressed as

Re = ρ u L / μ

= ρ u2 / (μ u / L)

=uL/ν (1)

where

Re = Reynolds Number (non-dimensional)

ρ = density (kg/m3, lbm/ft3 )

u = velocity based on the actual cross section area of the duct or pipe (m/s, ft/s)

μ = dynamic viscosity (Ns/m2, lbm/s ft)

L = characteristic length (m, ft)

ν = μ / ρ = kinematic viscosity (m2/s, ft2/s)

Reynolds Number for Flow in Pipe or Duct

For a pipe or duct the characteristic length is the hydraulic diameter.

L = dh

where

dh = hydraulic diameter (m, ft)

The Reynolds Number for the flow in a duct or pipe can with the hydraulic diameter be
expressed as

Re = ρ u dh / μ

= u dh / ν (2)

where

dh = hydraulic diameter (m, ft)

Reynolds Number for a Pipe or Duct in Imperial Units

The Reynolds number for a pipe or duct expressed in Imperial units


Re = 7745.8 u dh / ν (2a)

where

Re = Reynolds Number (non dimensional)

u = velocity (ft/s)

dh = hydraulic diameter (in)

ν = kinematic viscosity (cSt) (1 cSt = 10-6 m2/s )

The Reynolds Number can be used to determine if flow is laminar, transient or turbulent. The
flow is

 laminar - when Re < 2300


 transient - when 2300 < Re < 4000
 turbulent - when Re > 4000

In practice laminar flow is only actual for viscous fluids - like crude oil, fuel oil and other oils.

Example - Calculate Reynolds Number

A Newtonian fluid with a dynamic or absolute viscosity of 0.38 Ns/m2 and a specific gravity of
0.91 flows through a 25 mm diameter pipe with a velocity of 2.6 m/s.

Density can be calculated from the specific gravity of the fluid and the density of the specific
gravity reference water 1000 kg/m3 - as

ρ = 0.91 (1000 kg/m3)

= 910 kg/m3

Reynolds Number can then be calculated using equation (1) like

Re = (910 kg/m3) (2.6 m/s) (25 mm) (10-3 m/mm) / (0.38 Ns/m2)

= 156 ((kg m / s2)/N)

= 156 ~ Laminar flow

1 (N) = 1 (kg m / s2)

Related Mobile Apps from The Engineering ToolBox

 Reynolds Number - Calculator App


- free apps for offline use on mobile devices.

Online Reynolds Calculator

Density and absolute (dynamic) viscosity is Known

This calculator can be used if density and absolute (dynamic) viscosity of the fluid is known. The
calculator is valid for incompressible flow - flow with fluids or gases without compression - as
typical for air flows in HVAC systems or similar. The calculator is generic and can be used for
metric and imperial units as long as the use of units are consistent.

0.146

Density - ρ - (kg/m3, lbm/ft3)


20

Velocity - u - (m/s, ft/s)

0.5

Hydraulic diameter - dh - (or characteristic length - L) (m, ft)

0.0000122

Absolute (dynamic) viscosity - μ - (Ns/m2, lbm/s ft)

Default values are for air at 60 oF, 2 atm pressure and density 0.146 lbm/ft3, flowing 20 ft/s
between two metal sheets with characteristic length 0.5 ft. Dynamic (absolute) viscosity is 1.22
10-5 lbm/s ft.

Kinematic viscosity is known

The calculator below can be used when kinematic viscosity of the fluid is known. The calculator
is generic and can be used for metric and imperial units as long as the use of units are consistent.

Velocity - u - (m/s, ft/s)


0.102

Hydraulic diameter - dh - (or characteristic length - L) (m, ft)

0.000001004

Kinematic viscosity - ν - (m2/s, ft2/s)) (1 cSt = 10-6 m2/s)

Default values are for water at 20oC with kinematic viscosity 1.004 10-6 m2/s in a schedule 40
steel pipe. The characteristic length (or hydraulic diameter) of the pipe is 0.102 m.
Calculating Reynolds number in water tunnel

I am trying to calculate the Reynolds number of a flow that I will be creating in a water tunnel.
Reynolds number (RE) is given by the following where ρ gives fluid density, u gives velocity
of the fluid, L gives the characteristic linear dimension, and μ is the dynamic viscosity of the
fluid.
RE=ρuLμ

In my system, the flow velocity is 0.5

m/s. The density of water is 997 kg/m3 and the value I am using for the dynamic viscosity of
water is 8.9×10−4 Pa/s. The viewing area of the tunnel that my object will be in is a rectangle
with a square cross sectional area of 0.0225 m2. Plugging these values into the Reynold's
number equation, however, gives me a value of ≈84017

Based on this paper, I am skeptical that I used an incorrect value. The paper reports a more
powerful water tunnel with a significantly lower Reynolds number, 15000

Commercial water tunnels typically generate a momentum thickness based Reynolds number
(Reθ) ∼1000, which is slightly above the laminar to turbulent transition. The current work
compiles the literature on the design of high-Reynolds number facilities and uses it to design a
high-Reynolds number recirculating water tunnel that spans the range between commercial water
tunnels and the largest in the world. The final design has a 1.1 m long test-section with a 152 mm
square cross section that can reach speed of 10 m/s, which corresponds to Reθ=15,000. Flow
conditioning via a tandem configuration of honeycombs and settling-chambers combined with an
8.5:1 area contraction resulted in an average test-section inlet turbulence level <0.3% and
negligible mean shear in the test-section core. The developing boundary layer on the test-section
walls conform to a canonical zero-pressure-gradient (ZPG) flat-plate turbulent boundary layer
(TBL) with the outer variable scaled profile matching a 1/7th power-law fit, inner variable scaled
velocity profiles matching the log-law and a shape factor of 1.3.

I suspect that I may have used an incorrect value for L

. Could anyone advise me where I went wrong? I do believe that the flow around my object, a
turbine, is turbulent, however, I don't believe it is so much so as to having a Reynolds number of
80000.
Reynolds number
From Wikipedia, the free encyclopedia

Jump to navigation Jump to search

The plume from this candle flame goes from laminar to turbulent. The Reynolds number can be used to
predict where this transition will take place.

A vortex street around a cylinder. This can occur around cylinders and spheres, for any fluid, cylinder
size and fluid speed, provided that it has a Reynolds number between roughly 40 and 1000. [1]

George Stokes introduced Reynolds numbers.


Osborne Reynolds popularised the concept.

The Reynolds number (Re) is an important dimensionless quantity in fluid mechanics used to help
predict flow patterns in different fluid flow situations. At low Reynolds numbers, flows tend to be
dominated by laminar (sheet-like) flow, while at high Reynolds numbers 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 (eddy currents). These eddy currents begin to churn the flow, using up
energy in the process, which for liquids increases the chances of cavitation. The Reynolds number has
wide applications, ranging from liquid flow in a pipe to the passage of air over an aircraft wing. It is used
to predict the transition from laminar to turbulent flow, and is used in the scaling of similar but
different-sized flow situations, such as between an aircraft model in a wind tunnel and the full size
version. The predictions of the onset of turbulence and the ability to calculate scaling effects can be
used to help predict fluid behaviour on a larger scale, such as in local or global air or water movement
and thereby the associated meteorological and climatological effects.

The concept was introduced by George Stokes in 1851,[2] but the Reynolds number was named by Arnold
Sommerfeld in 1908[3] after Osborne Reynolds (1842–1912), who popularized its use in 1883.[4][5]

Contents
 1 Definition
 2 History
 3 Derivation
 4 Flow in a pipe
 5 Laminar–turbulent transition
 6 Flow in a wide duct
 7 Flow in an open channel
 8 Flow around airfoils
 9 Object in a fluid
o 9.1 In viscous fluids
o 9.2 Sphere in a fluid
o 9.3 Rectangular object in a fluid
o 9.4 Fall velocity
o 9.5 Packed bed
o 9.6 Stirred vessel
 10 Pipe friction
 11 Similarity of flows
 12 Smallest scales of turbulent motion
 13 In physiology
 14 Complex systems
 15 Derivation
 16 Relationship to other dimensionless parameters
 17 See also
 18 Notes
 19 References
 20 Further reading
 21 External links

Definition
The Reynolds number is the ratio of inertial forces to viscous forces within a fluid which is subjected to
relative internal movement due to different fluid velocities, which is known as a boundary layer in the
case of a bounding surface such as the interior of a pipe. A similar effect is created by the introduction of
a stream of high-velocity fluid into a low-velocity fluid, such as the hot gases emitted from a flame in air.
This relative movement generates fluid friction, which is a factor in developing turbulent flow.
Counteracting this effect is the viscosity of the fluid, which tends to inhibit turbulence. The Reynolds
number quantifies the relative importance of these two types of forces for given flow conditions, and is
a guide to when turbulent flow will occur in a particular situation.[6]

This ability to predict the onset of turbulent flow is an important design tool for equipment such as
piping systems or aircraft wings, but the Reynolds number is also used in scaling of fluid dynamics
problems, and is used to determine dynamic similitude between two different cases of fluid flow, such
as between a model aircraft, and its full-size version. Such scaling is not linear and the application of
Reynolds numbers to both situations allows scaling factors to be developed.

With respect to laminar and turbulent flow regimes:

 laminar flow occurs at low Reynolds numbers, where viscous forces are dominant, and is
characterized by smooth, constant fluid motion;
 turbulent flow occurs at high Reynolds numbers and is dominated by inertial forces, which tend
to produce chaotic eddies, vortices and other flow instabilities.[7]

The Reynolds number is defined as[3]

where:
 ρ is the density of the fluid (SI units: kg/m3)
 u is the velocity of the fluid with respect to the object (m/s)
 L is a characteristic linear dimension (m)
 μ is the dynamic viscosity of the fluid (Pa·s or N·s/m2 or kg/m·s)
 ν is the kinematic viscosity of the fluid (m2/s).

The Reynolds number can be defined for several different situations where a fluid is in relative motion to
a surface.[n 1] These definitions generally include the fluid properties of density and viscosity, plus a
velocity and a characteristic length or characteristic dimension (L in the above equation). This dimension
is a matter of convention – for example radius and diameter are equally valid to describe spheres or
circles, but one is chosen by convention. For aircraft or ships, the length or width can be used. For flow
in a pipe, or for a sphere moving in a fluid, the internal diameter is generally used today. Other shapes
such as rectangular pipes or non-spherical objects have an equivalent diameter defined. For fluids of
variable density such as compressible gases or fluids of variable viscosity such as non-Newtonian fluids,
special rules apply. The velocity may also be a matter of convention in some circumstances, notably
stirred vessels.

In practice, matching the Reynolds number is not on its own sufficient to guarantee similitude. Fluid flow
is generally chaotic, and very small changes to shape and surface roughness of bounding surfaces can
result in very different flows. Nevertheless, Reynolds numbers are a very important guide and are widely
used.

History

Osborne Reynolds's apparatus of 1883 demonstrating the onset of turbulent flow. The apparatus is still
at the University of Manchester.
Diagram from Reynolds's 1883 paper showing onset of turbulent flow.

Osborne Reynolds famously studied the conditions in which the flow of fluid in pipes transitioned from
laminar flow to turbulent flow. In his 1883 paper Reynolds described the transition from laminar to
turbulent flow in a classic experiment in which he examined the behaviour of water flow under different
flow velocities using a small stream of dyed water introduced into the centre of clear water flow in a
larger pipe.

The larger pipe was glass so the behaviour of the layer of the dyed stream could be observed, and at the
end of this pipe there was a flow control valve used to vary the water velocity inside the tube. When the
velocity was low, the dyed layer remained distinct through the entire length of the large tube. When the
velocity was increased, the layer broke up at a given point and diffused throughout the fluid's cross-
section. The point at which this happened was the transition point from laminar to turbulent flow.

From these experiments came the dimensionless Reynolds number for dynamic similarity—the ratio of
inertial forces to viscous forces. Reynolds also proposed what is now known as the Reynolds-averaging
of turbulent flows, where quantities such as velocity are expressed as the sum of mean and fluctuating
components. Such averaging allows for 'bulk' description of turbulent flow, for example using the
Reynolds-averaged Navier–Stokes equations.

Derivation
The form of the Reynolds number can be derived as follows[8]

where:

 is time
 is cross-sectional position

 is (local) flow speed

 is the shear stress (Pa)

 is the cross-sectional area of the flow

 is the volume of the fluid element

 is the maximum[9] speed of the object relative to the fluid (m/s)

 is a characteristic linear dimension (traveled length of the fluid; hydraulic diameter


when dealing with river systems) (m)

 is the dynamic viscosity of the fluid (Pa·s)

 is the kinematic viscosity ( ) (m2/s)

 is the density of the fluid (kg/m3).

Flow in a pipe
For flow in a pipe or tube, the Reynolds number is generally defined as[10]

where

DH is the hydraulic diameter of the pipe (the inside diameter if the pipe is circular) (m),

Q is the volumetric flow rate (m3/s),

A is the pipe's cross-sectional area (m2),

u is the mean velocity of the fluid (m/s),


μ (mu) is the dynamic viscosity of the fluid (Pa·s = N·s/m2 = kg/(m·s)),

ν (nu) is the kinematic viscosity (ν = μ/ρ ) (m2/s),

ρ is the density of the fluid (kg/m3).

For shapes such as squares, rectangular or annular ducts where the height and width are comparable,
the characteristic dimension for internal-flow situations is taken to be the hydraulic diameter, DH,
defined as

where A is the cross-sectional area, and P is the wetted perimeter. The wetted perimeter for a channel is
the total perimeter of all channel walls that are in contact with the flow. [11] This means that the length of
the channel exposed to air is not included in the wetted perimeter.

For a circular pipe, the hydraulic diameter is exactly equal to the inside pipe diameter:

For an annular duct, such as the outer channel in a tube-in-tube heat exchanger, the hydraulic diameter
can be shown algebraically to reduce to

where

Do is the inside diameter of the outer pipe,

Di is the outside diameter of the inner pipe.

For calculation involving flow in non-circular ducts, the hydraulic diameter can be substituted for the
diameter of a circular duct, with reasonable accuracy, if the aspect ratio AR of the duct cross-section
remains in the range 1/4 < AR < 4.[12]

Laminar–turbulent transition
Main article: Laminar–turbulent transition

In boundary layer flow over a flat plate, experiments confirm that, after a certain length of flow, a
laminar boundary layer will become unstable and turbulent. This instability occurs across different scales
and with different fluids, usually when Rex ≈ 5×105,[13] where x is the distance from the leading edge of
the flat plate, and the flow velocity is the freestream velocity of the fluid outside the boundary layer.
For flow in a pipe of diameter D, experimental observations show that for "fully developed" flow,[n 2]
laminar flow occurs when ReD < 2300 and turbulent flow occurs when ReD > 2900.[14][15] At the lower end
of this range, a continuous turbulent-flow will form, but only at a very long distance from the inlet of the
pipe. The flow in between will begin to transition from laminar to turbulent and then back to laminar at
irregular intervals, called intermittent flow. This is due to the different speeds and conditions of the fluid
in different areas of the pipe's cross-section, depending on other factors such as pipe roughness and
flow uniformity. Laminar flow tends to dominate in the fast-moving center of the pipe while slower-
moving turbulent flow dominates near the wall. As the Reynolds number increases, the continuous
turbulent-flow moves closer to the inlet and the intermittency in between increases, until the flow
becomes fully turbulent at ReD > 2900.[14] This result is generalized to non-circular channels using the
hydraulic diameter, allowing a transition Reynolds number to be calculated for other shapes of channel.
[14]

These transition Reynolds numbers are also called critical Reynolds numbers, and were studied by
Osborne Reynolds around 1895.[5] The critical Reynolds number is different for every geometry.[16]

Flow in a wide duct


For a fluid moving between two plane parallel surfaces—where the width is much greater than the
space between the plates—then the characteristic dimension is equal to twice the distance between the
plates. This is consistent with the annular duct and rectangular duct cases above taken to a limiting
aspect ratio.

Flow in an open channel


For flow of liquid with a free surface, the hydraulic radius must be determined. This is the cross-sectional
area of the channel divided by the wetted perimeter. For a semi-circular channel, it is quarter the
diameter (in case of full pipe flow). For a rectangular channel, the hydraulic radius is the cross-sectional
area divided by the wetted perimeter. Some texts then use a characteristic dimension that is four times
the hydraulic radius, chosen because it gives the same value of Re for the onset of turbulence as in pipe
flow,[17] while others use the hydraulic radius as the characteristic length-scale with consequently
different values of Re for transition and turbulent flow.

Flow around airfoils


Reynolds numbers are used in airfoil design to (among other things) manage "scale effect" when
computing/comparing characteristics (a tiny wing, scaled to be huge, will perform differently). [18] Fluid
dynamicists define the chord Reynolds number R like this: R = Vc/ν , where V is the flight speed, c is the
chord length, and ν is the kinematic viscosity of the fluid in which the airfoil operates, which is
1.460×10−5 m2/s for the atmosphere at sea level.[19] In some special studies a characteristic length other
than chord may be used; rare is the "span Reynolds number", which is not to be confused with spanwise
stations on a wing, where chord is still used.[20]

Object in a fluid
The Reynolds number for an object moving in a fluid, called the particle Reynolds number and often
denoted Rep, characterizes the nature of the surrounding flow and its fall velocity.

In viscous fluids

This section does not cite any sources. Please help improve this section by adding citations to
reliable sources. Unsourced material may be challenged and removed. (February 2014) (Learn
how and when to remove this template message)

The high viscosity of honey results in perfectly laminar flow when poured from a bucket, while the low
surface tension allows it to remain sheet-like even after reaching the fluid below. Analogous to
turbulence, when the flow meets resistance it slows and begins oscillating back and forth, piling upon
itself.
Creeping flow past a falling sphere: streamlines, drag force Fd and force by gravity Fg.

Where the viscosity is naturally high, such as polymer solutions and polymer melts, flow is normally
laminar. The Reynolds number is very small and Stokes' law can be used to measure the viscosity of the
fluid. Spheres are allowed to fall through the fluid and they reach the terminal velocity quickly, from
which the viscosity can be determined.

The laminar flow of polymer solutions is exploited by animals such as fish and dolphins, who exude
viscous solutions from their skin to aid flow over their bodies while swimming. It has been used in yacht
racing by owners who want to gain a speed advantage by pumping a polymer solution such as low
molecular weight polyoxyethylene in water, over the wetted surface of the hull.

It is, however, a problem for mixing of polymers, because turbulence is needed to distribute fine filler
(for example) through the material. Inventions such as the "cavity transfer mixer" have been developed
to produce multiple folds into a moving melt so as to improve mixing efficiency. The device can be fitted
onto extruders to aid mixing.

Sphere in a fluid

For a sphere in a fluid, the characteristic length-scale is the diameter of the sphere and the characteristic
velocity is that of the sphere relative to the fluid some distance away from the sphere, such that the
motion of the sphere does not disturb that reference parcel of fluid. The density and viscosity are those
belonging to the fluid.[21] Note that purely laminar flow only exists up to Re = 10 under this definition.

Under the condition of low Re, the relationship between force and speed of motion is given by Stokes'
law.[22]

Rectangular object in a fluid

The equation for a rectangular object is identical to that of a sphere, with the object being approximated
as an ellipsoid and the axis of length being chosen as the characteristic length scale. Such considerations
are important in natural streams, for example, where there are few perfectly spherical grains. For grains
in which measurement of each axis is impractical, sieve diameters are used instead as the characteristic
particle length-scale. Both approximations alter the values of the critical Reynolds number.

Fall velocity

The particle Reynolds number is important in determining the fall velocity of a particle. When the
particle Reynolds number indicates laminar flow, Stokes' law can be used to calculate its fall velocity.
When the particle Reynolds number indicates turbulent flow, a turbulent drag law must be constructed
to model the appropriate settling velocity.

Packed bed
For fluid flow through a bed, of approximately spherical particles of diameter D in contact, if the voidage
is ε and the superficial velocity is vs, the Reynolds number can be defined as[23]

or

or

The choice of equation depends on the system involved: the first is successful in correlating the data for
various types of packed and fluidized beds, the second Reynolds number suits for the liquid-phase data,
while the third was found successful in correlating the fluidized bed data, being first introduced for
liquid fluidized bed system.[23]

Laminar conditions apply up to Re = 10, fully turbulent from Re = 2000.[21]

Stirred vessel

In a cylindrical vessel stirred by a central rotating paddle, turbine or propeller, the characteristic
dimension is the diameter of the agitator D. The velocity V is ND where N is the rotational speed in rad
per second. Then the Reynolds number is:

The system is fully turbulent for values of Re above 10000.[24]

Pipe friction

The Moody diagram, which describes the Darcy–Weisbach friction factor f as a function of the Reynolds
number and relative pipe roughness.
Pressure drops[25] seen for fully developed flow of fluids through pipes can be predicted using the Moody
diagram which plots the Darcy–Weisbach friction factor f against Reynolds number Re and relative
roughness ε/D . The diagram clearly shows the laminar, transition, and turbulent flow regimes as
Reynolds number increases. The nature of pipe flow is strongly dependent on whether the flow is
laminar or turbulent.

Similarity of flows

Qualitative behaviors of fluid flow over a cylinder depends to a large extent on Reynolds number; similar
flow patterns often appear when the shape and Reynolds number is matched, although other
parameters like surface roughness have a big effect.

In order for two flows to be similar, they must have the same geometry and equal Reynolds and Euler
numbers. When comparing fluid behavior at corresponding points in a model and a full-scale flow, the
following holds:

where is the Reynolds number for the model, and is full-scale Reynolds number, and
similarly for the Euler numbers.

The model numbers and design numbers should be in the same proportion, hence
This allows engineers to perform experiments with reduced scale models in water channels or wind
tunnels and correlate the data to the actual flows, saving on costs during experimentation and on lab
time. Note that true dynamic similitude may require matching other dimensionless numbers as well,
such as the Mach number used in compressible flows, or the Froude number that governs open-channel
flows. Some flows involve more dimensionless parameters than can be practically satisfied with the
available apparatus and fluids, so one is forced to decide which parameters are most important. For
experimental flow modeling to be useful, it requires a fair amount of experience and judgment of the
engineer.

An example where the mere Reynolds number is not sufficient for similarity of flows (or even the flow
regime – laminar or turbulent) are bounded flows, i.e. flows that are restricted by walls or other
boundaries. A classical example of this is the Taylor–Couette flow, where the dimensionless ratio of radii
of bounding cylinders is also important, and many technical applications where these distinctions play
an important role.[26][27] Principles of these restrictions were developed by Maurice Marie Alfred Couette
and Geoffrey Ingram Taylor and developed further by Floris Takens and David Ruelle.

Typical values of Reynolds number[28][29]

 Bacterium ~ 1 × 10−4
 Ciliate ~ 1 × 10−1
 Smallest fish ~ 1
 Blood flow in brain ~ 1 × 102
 Blood flow in aorta ~ 1 × 103
 Onset of turbulent flow ~ 2.3 × 103 to 5.0 × 104 for pipe flow to 106 for boundary layers
 Typical pitch in Major League Baseball ~ 2 × 105
 Person swimming ~ 4 × 106
 Fastest fish ~ 1 × 108
 Blue whale ~ 4 × 108
 A large ship (RMS Queen Elizabeth 2) ~ 5 × 109

Smallest scales of turbulent motion


This section does not cite any sources. Please help improve this section by adding citations to
reliable sources. Unsourced material may be challenged and removed. (January 2019) (Learn how
and when to remove this template message)

In a turbulent flow, there is a range of scales of the time-varying fluid motion. The size of the largest
scales of fluid motion (sometimes called eddies) are set by the overall geometry of the flow. For
instance, in an industrial smoke stack, the largest scales of fluid motion are as big as the diameter of the
stack itself. The size of the smallest scales is set by the Reynolds number. As the Reynolds number
increases, smaller and smaller scales of the flow are visible. In a smoke stack, the smoke may appear to
have many very small velocity perturbations or eddies, in addition to large bulky eddies. In this sense,
the Reynolds number is an indicator of the range of scales in the flow. The higher the Reynolds number,
the greater the range of scales. The largest eddies will always be the same size; the smallest eddies are
determined by the Reynolds number.

What is the explanation for this phenomenon? A large Reynolds number indicates that viscous forces
are not important at large scales of the flow. With a strong predominance of inertial forces over viscous
forces, the largest scales of fluid motion are undamped—there is not enough viscosity to dissipate their
motions. The kinetic energy must "cascade" from these large scales to progressively smaller scales until
a level is reached for which the scale is small enough for viscosity to become important (that is, viscous
forces become of the order of inertial ones). It is at these small scales where the dissipation of energy by
viscous action finally takes place. The Reynolds number indicates at what scale this viscous dissipation
occurs.

In physiology
This section does not cite any sources. Please help improve this section by adding citations to
reliable sources. Unsourced material may be challenged and removed. (December 2014) (Learn
how and when to remove this template message)

Poiseuille's law on blood circulation in the body is dependent on laminar flow. In turbulent flow the flow
rate is proportional to the square root of the pressure gradient, as opposed to its direct proportionality
to pressure gradient in laminar flow.

Using the definition of the Reynolds number we can see that a large diameter with rapid flow, where the
density of the blood is high, tends towards turbulence. Rapid changes in vessel diameter may lead to
turbulent flow, for instance when a narrower vessel widens to a larger one. Furthermore, a bulge of
atheroma may be the cause of turbulent flow, where audible turbulence may be detected with a
stethoscope.

Complex systems
Reynolds number interpretation has been extended into the area of arbitrary complex systems. Such as
financial flows,[30] nonlinear networks,[31] etc. In the latter case an artificial viscosity is reduced to
nonlinear mechanism of energy distribution in complex network media. Reynolds number then
represents a basic control parameter which expresses a balance between injected and dissipated energy
flows for open boundary system. It has been shown [31] that Reynolds critical regime separates two types
of phase space motion: accelerator (attractor) and decelerator. High Reynolds number leads to a chaotic
regime transition only in frame of strange attractor model.

Derivation
This section does not cite any sources. Please help improve this section by adding citations to
reliable sources. Unsourced material may be challenged and removed. (June 2013) (Learn how
and when to remove this template message)

The Reynolds number can be obtained when one uses the nondimensional form of the incompressible
Navier–Stokes equations for a newtonian fluid expressed in terms of the Lagrangian derivative:

Each term in the above equation has the units of a "body force" (force per unit volume) with the same
dimensions of a density times an acceleration. Each term is thus dependent on the exact measurements
of a flow. When one renders the equation nondimensional, that is when we multiply it by a factor with
inverse units of the base equation, we obtain a form that does not depend directly on the physical sizes.
One possible way to obtain a nondimensional equation is to multiply the whole equation by the factor

where

V is the mean velocity, v or v, relative to the fluid (m/s),

L is the characteristic length (m),

ρ is the fluid density (kg/m3).

If we now set

we can rewrite the Navier–Stokes equation without dimensions:

where the term μ/ρLV = 1/Re .


Finally, dropping the primes for ease of reading:

This is why mathematically all Newtonian, incompressible flows with the same Reynolds number are
comparable. Notice also that in the above equation, the viscous terms vanish for Re → ∞. Thus flows
with high Reynolds numbers are approximately inviscid in the free stream.
Relationship to other dimensionless parameters
There are many dimensionless numbers in fluid mechanics. The Reynolds number measures the ratio of
advection and diffusion effects on structures in the velocity field, and is therefore closely related to
Péclet numbers, which measure the ratio of these effects on other fields carried by the flow, for example
temperature and magnetic fields. Replacement of the kinematic viscosity ν = μ/ρ in Re by the thermal or
magnetic diffusivity results in respectively the thermal Péclet number and the magnetic Reynolds
number. These are therefore related to Re by products with ratios of diffusivities, namely the Prandtl
number and magnetic Prandtl number.

See also
 Reynolds transport theorem
 Drag coefficient – Dimensionless parameter to quantify fluid resistance
 Deposition (geology) – Geological process in which sediments, soil and rocks are added to a
landform or land mass

Notes
1.

 The definition of the Reynolds number is not to be confused with the Reynolds equation or lubrication
equation.

 Full development of the flow occurs as the flow enters the pipe, the boundary layer thickens and then
stabilizes after several diameters distance into the pipe.

The dimensionless Reynolds number plays a prominent role in foreseeing the patterns in a fluid’s
behavior. The Reynolds number, referred to as Re, is used to determine whether the fluid flow is
laminar or turbulent. It is one of the main controlling parameters in all viscous flows where a
numerical model is selected according to pre-calculated Reynolds number.

Although the Reynolds number comprises both static and kinematic properties of fluids, it is
specified as a flow property since dynamic conditions are investigated. Technically speaking, the
Reynolds number is the ratio of the inertial forces and the viscous forces. In practice, the
Reynolds number is used to predict if the flow will be laminar or turbulent.
If the inertial forces, which resist a change in velocity of an object and are the cause of the fluid
movement, are dominant, the flow is turbulent. Otherwise, if the viscous forces, defined as the
resistance to flow, are dominant – the flow is laminar. The Reynolds number can be specified as
below:

Re=inertial forceviscous force=fluid and flow propertiesfluid properties(


1)

For instance, a glass of water which stands on a static surface, regardless of any forces apart
from gravity, is at rest and flow properties are ignored. Thus, the numerator in equation (1) is
“0”. That results in independence from the Reynolds number for a fluid at rest. On the other
hand, whilst water is spilled by tilting a water-filled glass, flow properties abide by physical
laws, a Reynolds number can be estimated as to predict fluid flow that is illustrated in Figure 1.

Figure 1: A glass of water which is a) at rest; b) flows

History
The theory of a dimensionless number which predicts fluid flow was initially introduced by Sir
George Stokes (1819-1903) who had attempted to figure out drag force on a sphere whereas
neglecting the inertial term. Stokes had also carried out the studies of Claude Louis Navier
(1785-1836) taking them further and deriving the equation of motion by adding a viscous term in
1851 – thereby revealing the Navier-Stokes equation [1].

Stokes flow, named after Stokes’ approach to viscous fluid flow, is the mathematical model in
which the Reynolds number is so low that it is presumed to be zero. Various scientists had
conducted studies to examine properties of fluid movement after Stokes. Even though the
Navier-Stokes equations thoroughly analyzed fluid flow, it was quite hard to apply them for
arbitrary flows where the Reynolds number could easily predict fluid movement.

In 1883, Irish scientist Osborne Reynolds discovered the dimensionless number that predicts
fluid flow based on static and dynamic properties such as velocity, density, dynamic viscosity
and characteristics of the fluid [2]. He conducted experimental studies to examine the
relationship between the velocity and behavior of fluid flow. For this purpose, an experimental
setup (Figure 2.a) has been established by Reynolds using dyed water which was released in the
middle of the cross-sectional area into the main clear water to visualize the movement of fluid
flow through the glass tube (Figure 2.b).

Figure 2: a) Experimental setup established by Osborne Reynolds; b) Experimental visualization


of laminar and turbulent flow

The study of Osborne Reynolds titled ‘an experimental investigation of the circumstances which
determine whether the motion of water in parallel channels shall be direct or sinuous’ regarding
the dimensionless number was issued in “Philosophical Transactions of the Royal Society”.
According to the article, the dimensionless number discovered by Reynolds was suitable to
foresee fluid flow in a broad range from water flow in a pipe to airflow over an airfoil.[2]
Figure 3: Osborne Reynolds (1842-1912)

The dimensionless number was referred to as parameter :math:‘R’, until the presentation of
German physicist Arnold Sommerfeld (1868 – 1951) at the 4th International Congress of
Mathematicians in Rome (1908), where he referred to the ‘R’ number as the ‘Reynolds number’.
The term used by Sommerfeld has been used worldwide ever since. [3]

Derivation
The dimensionless Reynolds number predicts whether the fluid flow would be laminar or
turbulent referring to several properties such as velocity, length, viscosity, and also type of flow.
It is expressed as the ratio of inertial forces to viscous forces and can be explained in terms of
units and parameters respectively, as below:

Re=ρVLμ=VLv(2)
Re=FinertiaFviscous=kgm3×ms×mPa×s=FF(3)

where ρ (kgm3)

is the density of the fluid, V (ms2) is the characteristic velocity of the flow, and L (m) is the
characteristic length scale of flow. (4) Equation (3) is the derivation of units at which the
Reynolds number is specified as non-dimensional. Variations of the Reynolds number are shown
in equation (2) where μ(Pa×s) is the dynamic viscosity of fluid and v (m2s)

is the kinematic viscosity. The transition between dynamic and kinematic viscosity is as follows:

v=μρ(4)

Fluid, Flow and Reynolds Number


The applicability of the Reynolds number differs depending on the specifications of the fluid
flow such as the variation of density (compressibility), variation of viscosity (Non-Newtonian),
being internal- or external flow etc. The critical Reynolds number is the expression of the value
to specify transition among regimes which diversifies regarding type of flow and geometry as
well. Whilst the critical Reynolds number for turbulent flow in a pipe is 2000, the critical
Reynolds number for turbulent flow over a flat plate, when the flow velocity is the free-stream
velocity, is in a range from 105

to 106. 4

The Reynolds number also predicts the viscous behavior of the flow in case fluids are
Newtonian. Therefore, it is highly important to perceive the physical case to avoid inaccurate
predictions. Transition regimes and internal as well as external flows with either low or high
Reynolds number in use, are the basic fields to comprehensively investigate the Reynolds
number. Newtonian fluids are fluids that have a constant viscosity. If the temperature stays the
same, it does not matter how much stress is applied on a Newtonian fluid; it will always have the
same viscosity. Examples include water, alcohol and mineral oil.

Laminar to turbulent transition


The fluid flow can be specified under two different regimes: Laminar and Turbulent. The
transition among the regimes is an important issue that is driven by both fluid and flow
properties. As mentioned before, the critical Reynolds number, which changes in accordance
with the physical case, can be classified as internal and external, where it might face slight
changes in the amount. Yet while the Reynolds number regarding the laminar-turbulent
transition can be defined reasonably for internal flow, it is hard to specify a definition for an
external flow.

Internal flow
The fluid flow in a pipe as an internal flow had been illustrated by Reynolds as in Figure 2.b.
The critical Reynolds number for internal flow is: [4]

Flow type Renolds Number Range


Laminar regime up to Re=2300
Transition regime 2300<Re<4000
Turbulent regime Re>4000

Table 1: Different Reynolds Numbers for different types of flow

Open-channel flow, fluid flow in an object and flow with pipe friction are internal flows in
which the Reynolds number is predicted based on hydraulic diameter D

instead of characteristic length L. In case the pipe is cylindrical, the hydraulic diameter D
is accepted as the actual diameter of the cylinder, meaning the Reynolds number is as follows:

Re=FinertiaFviscous=ρVDHμ(5)

The shape of a pipe or duct can vary (e.g. square, rectangular, etc.). In those cases, the hydraulic
diameter is determined as below:

DH=4AP(6)

where A is the cross-sectional area and P is the wetted perimeter.

The friction on the pipe surface due to roughness is an effective parameter to consider because it
causes laminar to turbulence transition and energy losses. The ‘Moody Chart’ (Figure 4) was
generated by Lewis Ferry Moody (1944) to predict fluid flow in pipes where roughness was
effective. It is a practical method to determine energy losses in terms of friction factor due to
roughness throughout the inner surface of a pipe. The critical Reynolds number for a pipe with
surface roughness abides by the regimes above [2]. In the chart below you can see a logarithmic
scale at the bottom with a scale for the friction factor at the left and the relative roughness of the
pipe at the right.
Figure 4: The Moody chart for pipe friction with smooth and rough walls

External flow
External flow at which mainstream has no district boundaries is alike to internal flow that also
has a transition regime. Flows over bodies such as a flat plate, cylinder and sphere are the
standard cases used to investigate the effect of velocity throughout the stream. In 1914, German
scientist Ludwig Prandtl discovered the boundary layer, which is partially the function of the
Reynolds number, covering surface through laminar, turbulent and also transition regimes [5].
The flow over a flat surface is shown in figure 4 with regimes where xc

is the critical length for transition, L is total length of the plate and u

is the velocity of the free stream flow.

Figure 5: Transition of boundaries through the flow over the flat plate surface as an example of
external flow

In general, the boundary layer dilates with movement through x

direction (any point on plate throughout L


) that eventually results in unstable conditions where the Reynolds number increases
simultaneously. The critical Reynolds number for flow over flat plate surface is:

Recritical=ρVxμ≥3×105 to 3×106(7)

which depends on the uniformity of the flow over the surface. Yet while the critical Reynolds
numbers for regimes are virtually specified for internal flow, it is hard to detect them for external
flow which diversifies the critical Reynolds number regarding geometry. Furthermore, apart
from internal flow, boundary layer separation is an anomalous issue for external flow where
several ambiguities are encountered to generate a reliable numerical model with respect to a
physical domain.[6]

Low and high Reynolds number


The Reynolds number, the ratio of inertial and viscous effects, is also effective on Navier-Stokes
equations to truncate mathematical models. While Re→∞

, the viscous effects are presumed negligible where viscous terms in Navier-Stokes equations are
dropped. The simplified form of the Navier-Stokes equations — called Euler equations — can be
specified as follows:

DρDt=−ρ∇×u(8)
DuDt=−∇pρ+g(9)
DeDt=−pρ∇×u(10)

where ρ

is density, u is velocity, p is pressure, g is gravitational acceleration, and e

is the specific internal energy.[6] Though viscous effects are relatively important for fluids, the
inviscid flow model partially provides a reliable mathematical model as to predict a real process
for some specific cases. For instance, high-speed external flow over bodies is a broadly used
approximation where the inviscid approach fits reasonably.

While Re≪1

, the inertial effects are presumed negligible and related terms in the Navier-Stokes equations can
be dropped. The simplified form of Navier-Stokes equations is called either Creeping or Stokes
flow:

μ∇2u−∇p+f=0(11)
∇×u=0(12)

where u
is the velocity of the fluid, ∇p is the pressure gradient, μ is the dynamic viscosity and f

is the applied body force. [6] Having tangible viscous effects, the creeping flow is a suitable
approach that can be used to investigate e.g. the flow of lava, swimming of microorganisms,
flow of polymers, lubrication, etc.

Application of the Reynolds number


The numerical solution of fluid flow relies on mathematical models which have been generated
by both experimental studies and related physical laws. One of the significant steps throughout
the numerical examination is to determine an appropriate mathematical model that simulates the
physical domain. To obtain a reasonably good prediction for the behavior of fluids under various
circumstances, the Reynolds number has been accepted as a substantial prerequisite for fluid
flow analysis. For instance, movement of glycerin in a circular duct can be predicted by the
Reynolds number as follows: [7]

Matter Glycerin

Dencity at 23∘C
1259
Dynamic Viscosity
(Pa.s)
0.950
Diameter of Duct
(m)
0.05
Velocity of Glycerin flow at inlet
(m/s)
0.5
ReGlycerin=ρVDHμ=1259×0.5×0.050.950≈33.1(13)

where glycerin flow is laminar in accordance with the critical Reynolds number for internal flow.

Reynolds number & SimScale


The Reynolds number is never really visible in SimScale’s simulation projects but it influences
many of them. Here are some interesting blog posts to read about the Reynolds number:

What Everybody Ought to Know About CFD

How Dimples on a Golf Ball Affect Its Flight and Aerodynamics


10 Piping Design Simulations: Fluid Flow and Stress Analyses

Fluid Viscosity Properties

Fluid Viscosity

Fluid Viscosity, sometimes referred to as dynamic viscosity or absolute viscosity, is the fluid's
resistance to flow, which is caused by a shearing stress within a flowing fluid and between a flowing fluid
and its container.

Viscosity is usually denoted by the Greek symbol μ (mu) and is defined as the ratio of shearing stress τ
(Greek letter tau) to the rate of change of velocity, v, which in mathematical terms can be expressed as
dv/dy (where this is the derivative of the of the velocity with respect to the distance y).

The derivative dv/dy is called the velocity gradient.

This results in the important equation for fluid shear for viscous or laminar flow:

τ = μ•dv/dy

However, the above equation is not applicable for turbulent flow where a large amount of the shear
stress is due to the exchange of momentum between adjacent layers of the fluid. To determine whether
the flow is laminar or turbulent requires that you calculate the Reynold's number of the flowing fluid.
Laminar flow will occur where the Reynold's number is less than 2300.

From the above equation it can be determined that the dimensions of viscosity are force multiplied by
time divided by length squared or FT/L² The units of viscosity in the English system and the SI system
are:
lb•sec/ft² or Slug/ft•sec and N•s/m² or kg/m•s

Dynamic Viscosity / Absolute Viscosity

The Pascal unit (Pa) is specifies pressure, or stress = force per area

Pascals can be combined with time (seconds) to define dynamic viscosity.

μ = Pa•s
1.00 Pa•s = 10 Poise = 1000 Centipoise

Centipoise (cP) is commonly used to describe dynamic viscosity because water at a temperature of 20°C
has a viscosity of 1.002 Centipoise.

This value must be converted back to 1.002 x 10^-3 Pa•s for use in calculations.

Kinematic Viscosity

Viscosity can be measured by timing the flow of a known volume of fluid from a viscosity measuring cup.
The timings can be used in a formula to estimate the kinematic viscosity value of the fluid in Centistokes
(cSt).

The motive force driving the fluid out of the cup is the head of fluid, which is also contained within the
equation that makes up the volume of the fluid. When the equations are rationalized the fluid head
term is eliminated leaving the units of Kinematic viscosity as area / time.

v = m²/s

1.0 m²/s = 10000 Stokes = 1000000 Centistokes

Water at a temperature of 20°C has a viscosity of 1.004 x 10^-6 m²/s


This evaluates to 1.004000 Centistokes.
This value must be converted back to 1.004 x 10^-6 m²/s for use in calculations.

The kinematic viscosity can also be determined by dividing the dynamic viscosity by the fluid density.

Kinematic Viscosity and Dynamic Viscosity Relationship

Kinematic Viscosity = Dynamic Viscosity / Density

v = μ/ρ
Dynamic, absolute and kinematic viscosities - convert
between CentiStokes (cSt), centipoises (cP), Saybolt Universal
Seconds (SSU) and degree Engler
Sponsored Links

Viscosity is an important fluid property when analyzing liquid behavior and fluid motion near
solid boundaries. The viscosity of a fluid is a measure of its resistance to gradual deformation by
shear stress or tensile stress. The shear resistance in a fluid is caused by inter-molecular friction
exerted when layers of fluid attempt to slide by one another.

 viscosity is the measure of a fluid's resistance to flow

 molasses is highly viscous


 water is medium viscous
 gas is low viscous

There are two related measures of fluid viscosity

 dynamic (or absolute)


 kinematic

Dynamic (absolute) Viscosity


Absolute viscosity - coefficient of absolute viscosity - is a measure of internal resistance.
Dynamic (absolute) viscosity is the tangential force per unit area required to move one horizontal
plane with respect to an other plane - at an unit velocity - when maintaining an unit distance
apart in the fluid.

The shearing stress between the layers of a non turbulent fluid moving in straight parallel lines
can be defined for a Newtonian fluid as

Shear stress can be expressed

τ = μ dc / dy

=μγ (1)

where

τ = shearing stress in fluid (N/m2)

μ = dynamic viscosity of fluid (N s/m2)

dc = unit velocity (m/s)

dy = unit distance between layers (m)

γ = dc / dy = shear rate (s-1)

Equation (1) is known as the Newtons Law of Friction.

(1) can be rearranged to express Dynamic viscosity as

μ = τ dy / dc

=τ /γ (1b)
In the SI system the dynamic viscosity units are N s/m2, Pa s or kg/(m s) - where

 1 Pa s = 1 N s/m2 = 1 kg/(m s)

Dynamic viscosity may also be expressed in the metric CGS (centimeter-gram-second) system as
g/(cm s), dyne s/cm2 or poise (p) where

 1 poise = 1 dyne s/cm2 = 1 g/(cm s) = 1/10 Pa s = 1/10 N s/m2

For practical use the Poise is normally too large and the unit is therefore often divided by 100 -
into the smaller unit centipoise (cP) - where

 1 P = 100 cP
 1 cP = 0.01 poise = 0.01 gram per cm second = 0.001 Pascal second = 1 milliPascal
second = 0.001 N s/m2

Water at 20.2oC (68.4oF) has the absolute viscosity of one - 1 - centiPoise.

Absolute Viscosity *)
Liquid
(N s/m2, Pa s)
Air 1.983 10-5
Water 10-3
Olive Oil 10-1
Glycerol 100
Liquid Honey 101
Golden Syrup 102
Glass 1040

*) at room temperature

 Absolute or Dynamic Viscosity of common Liquids

Kinematic Viscosity

Kinematic viscosity is the ratio of - absolute (or dynamic) viscosity to density - a quantity in
which no force is involved. Kinematic viscosity can be obtained by dividing the absolute
viscosity of a fluid with the fluid mass density like

ν=μ/ρ (2)

where

ν = kinematic viscosity (m2/s)

μ = absolute or dynamic viscosity (N s/m2)


ρ = density (kg/m3)

In the SI-system the theoretical unit of kinematic viscosity is m2/s - or the commonly used Stoke
(St) where

 1 St (Stokes) = 10-4 m2/s = 1 cm2/s

Stoke comes from the CGS (Centimetre Gram Second) unit system.

Since the Stoke is a large unit it is often divided by 100 into the smaller unit centiStoke (cSt) -
where

 1 St = 100 cSt
 1 cSt (centiStoke) = 10-6 m2/s = 1 mm2/s
 1 m2/s = 106 centiStokes

The specific gravity for water at 20.2oC (68.4oF) is almost one, and the kinematic viscosity for
water at 20.2oC (68.4oF) is for practical purpose 1.0 mm2/s (cStokes). A more exact kinematic
viscosity for water at 20.2oC (68.4oF) is 1.0038 mm2/s (cSt).

 Kinematic viscosity of common liquids and fluids

A conversion from absolute to kinematic viscosity in Imperial units can be expressed as

ν = 6.7197 10-4 μ / γ (2a)

where

ν = kinematic viscosity (ft2/s)

μ = absolute or dynamic viscosity (cP)

γ = specific weight (lb/ft3)

Viscosity and Reference Temperature

The viscosity of a fluid is highly temperature dependent - and for dynamic or kinematic viscosity
to be meaningful the reference temperature must be quoted. In ISO 8217 the reference
temperature for a residual fluid is 100oC. For a distillate fluid the reference temperature is 40oC.

 for a liquid - the kinematic viscosity decreases with higher temperature


 for a gas - the kinematic viscosity increases with higher temperature

Related Mobile Apps from The Engineering ToolBox

 Kinematic viscosity converter App


This is a free app that can be used offline on mobile devices.

Other Viscosity Units

Saybolt Universal Seconds (or SUS, SSU)

Saybolt Universal Seconds (or SUS) is an alternative unit for measuring viscosity. The efflux
time is Saybolt Universal Seconds (SUS) required for 60 milliliters of a petroleum product to
flow through the calibrated orifice of a Saybolt Universal viscometer - under a carefully
controlled temperature and as prescribed by test method ASTM D 88. This method has largely
been replaced by the kinematic viscosity method. Saybolt Universal Seconds is also called the
SSU number (Seconds Saybolt Universal) or SSF number (Saybolt Seconds Furol).

Kinematic viscosity in SSU versus dynamic or absolute viscosity can be expressed as

νSSU = B μ / SG

= B νcentiStokes (3)

where

νSSU = kinematic viscosity (SSU)

B = 4.632 for temperature 100 oF (37.8 oC)

B = 4.664 for temperature 210oF (98.9 oC)

μ = dynamic or absolute viscosity (cP)


SG = Specific Gravity
νcentiStokes = kinematic viscosity (centiStokes)

 SSU at other temperatures

Degree Engler

Degree Engler is used in Great Britain as a scale to measure kinematic viscosity. Unlike the
Saybolt and Redwood scales, the Engler scale is based on comparing the flow of the substance
being tested to the flow of another substance - water. Viscosity in Engler degrees is the ratio of
the time of a flow of 200 cubic centimeters of the fluid whose viscosity is being measured - to
the time of flow of 200 cubic centimeters of water at the same temperature (usually 20oC but
sometimes 50oC or 100oC) in a standardized Engler viscosity meter.

Newtonian Fluids
A fluid where the shearing stress is linearly related to the rate of shearing strain - is designated as
a Newtonian Fluid.

A Newtonian material is referred to as true liquid since the viscosity or consistency is not
affected by shear such as agitation or pumping at a constant temperature. Most common fluids -
both liquids and gases - are Newtonian fluids. Water and oils are examples of Newtonian liquids.

Shear-thinning or Pseudo-plastic Fluids

A Shear-thinning or pseudo-plastic fluid is a fluid where the viscosity decrease with increased
shear rate. The structure is time-independent.

Thixotropic Fluids

A Thixotropic fluid has a time-dependent structure. The viscosity of a thixotropic fluid decreases
with increasing time - at a constant shear rate.

Ketchup and mayonnaise are examples of thixotropic materials. They appear thick or viscous but
are possible to pump quite easily.

Dilatant Fluids

A Shear Thickening Fluid - or Dilatant Fluid - increases the viscosity with agitation or shear
strain. Dilatant fluids are known as non-Newton fluids.

Some dilatant fluids can become almost solid in a pump or pipe line. With agitation cream
becomes butter and candy compounds. Clay slurry and similar heavily filled liquids do the same
thing.

Bingham Plastic Fluids

A Bingham Plastic Fluid has a yield value which must be exceeded before it will start to flow
like a fluid. From that point the viscosity decreases with increasing agitation. Toothpaste,
mayonnaise and tomato ketchup are examples of such products.

Example - Air, Convert between Kinematic and Absolute Viscosity

Kinematic viscosity of air at 1 bar (1 105 Pa, N/m2) and 40oC is 16.97 cSt (16.97 10-6 m2/s).

The density of the air can be estimated with the Ideal Gas Law

ρ = p / (R T)

= (1 105 N/m2) / ( (287 J/(kg K)) ((273 oC) + (33 oC)) )

= 1.113 (kg/m3)
where

ρ = density (kg/m3)

p = absolute pressure (Pa, N/m2)

R = individual gas constant (J/(kg K))

T = absolute temperature (K)

The absolute viscosity can be calculated as

μ = 1.113 (kg/m3) 16.97 10-6 (m2/s)

= 1.88 10-5 (kg/(m s), N s/m2)

Viscosity of some Common Liquids

Saybolt Second
centiStokes
Universal Typical liquid
(cSt, 10-6 m2/s, mm2/s)
(SSU, SUS)
0.1 Mercury
1 31 Water (20oC)
Milk
4.3 40 SAE 20 Crankcase Oil
SAE 75 Gear Oil
15.7 80 No. 4 fuel oil
20.6 100 Cream
43.2 200 Vegetable oil
SAE 30 Crankcase Oil
110 500
SAE 85 Gear Oil
Tomato Juice
220 1000 SAE 50 Crankcase Oil
SAE 90 Gear Oil
440 2000 SAE 140 Gear Oil
Glycerine (20oC)
1100 5000
SAE 250 Gear Oil
2200 10000 Honey
6250 28000 Mayonnaise
19000 86000 Sour cream

 Viscosity Converting Chart

Kinematic viscosity can be converted from SSU to Centistokes with


νCentistokes = 0.226 νSSU - 195 / νSSU (4)

where

νSSU < 100

νCentistokes = 0.220 νSSU - 135 / νSSU

where

νSSU > 100

Viscosity and Temperature

Kinematic viscosity of fluids like water, mercury, oils SAE 10 and oil no. 3 - and gases like air,
hydrogen and helium are indicated in the diagram below. Note that

 for liquids - viscosity decreases with temperature


 for gases - viscosity increases with temperature
Measuring Viscosity

Three types of devices are used to measure viscosity

 capillary tube viscometer


 Saybolt viscometer
 rotating viscometer
Learn more about Dynamic Viscosity
Gas Turbine Fuel Systems and Fuels
Claire Soares, in Gas Turbines (Second Edition), 2015

Viscosity, Dynamic and Kinematic

is the resistance to movement of one layer of a fluid over another and is defined by Formula
F7.8. Kinematic viscosity is dynamic viscosity divided by density (Formula F7.9) and is the ratio
of viscous forces to inertia forces. Dynamic viscosity for liquid fuels is occasionally required for
performance calculations in determining fuel pump power requirements such as during start
modeling. Kinematic viscosity is a second-order variable in the calibration of bulk meters and
turbine flow meters for measuring liquid fuel volumetric flow rate.

Figure 7–23 presents kinematic viscosity for gas turbine liquid fuels. It is common practice to
arrange fuel dynamic viscosity as an input value to a performance computer program. Hence
suitable values may be taken from Figures 7–22 and 7–23. If necessary the reader may “fit” a
polynomial to the data presented in these charts to embed the properties within the computer
program.

Sign in to download full-size image

FIGURE 7–23. Oil viscosity versus temperature.


(Source: Rolls Royce.)

Liquid fuels are not “pumpable” with a kinematic viscosity of less than 1 cSt, and atomization
will be unsatisfactory above 10 cSt. Figure 7–22 shows that kerosenes generally have a viscosity
of less than 10 cSt even at –50°C and hence are always acceptable for atomization. However,
diesels do exceed the limit and hence fuel heating is required if its temperature is allowed to fall
below the threshold level.

None of the above effects are of importance for gas fuel as its viscosity is an order of magnitude
lower; hence a database is not provided here.

Read full chapter

Surface Engineering Concepts


David A. Simpson P.E., in Practical Onshore Gas Field Engineering, 2017

4.1.1.1 Dynamic viscosity (μ)

Dynamic viscosity can be determined via an apparatus that has concentric tubes whose annular
space can be filled with the fluid under test. One of the tubes is rotated and the amount of force
that it takes to rotate at a fixed angular velocity is proportional to the dynamic viscosity. It can
also be predicted through an equation of state or one of many empirical relationships. The
various methods can vary by more than 20% one from the next for the same fluid. None of the
empirical methods provide results that match field conditions when there is any acid gas in a gas
stream, and the equations of state methods are not a lot better for acid gases. With the wide range
of answers that you get from repeated lab measurement and/or using various equations, it is
generally best to pick a single method and pretend that you believe it—repeatable results will not
be available for viscosity measurements.

The basic units of dynamic viscosity are Poise (poise), Centipoise (cP), lbm/ft/s, or Pa×s. The
relationship between various units is given in Table 4.1.

Table 4.1. Relationship between viscosity units

Poise cP lbm/ft/s Pa×s


poise 1 0.001 0.0672 0.1
cP 100 1 0.000672 0.001
lbm/ft/s 14.88 1488 1 1.488
Pa×s 10 1000 0.672 1

Gas viscosity tends to be in the range of 0.01–0.02 cP (10×10−6 to 20×10−6 Pa×s) (air at 60°F
(15.6°C) is usually reported as 0.0179 cP (17.9×10−6 Pa×s)). Liquid viscosities tend to be much
higher (water at 60°F (15.6°C) is usually reported as 1 cP (0.001 Pa×s)). Dynamic viscosity is a
strong function of temperature and a weak function of pressure.
Read full chapter

Hydrodynamic Lubrication
In Tribology Series, 1990

2.9 Viscosity of Gases

The dynamic viscosity of gases is very small and does not change much from one gas to another,
under normal temperature and pressure, as seen in Table 2.10.

Table 2.10. Dynamic viscosity of gases.

Gas Hydrogen Neon Nitrogen Air


Viscosity Pa.s 9.10−6 31.10−6 18.10−6 18.10−6

The kinematic viscosity of gases is higher since their density is small. Thus for air, under normal
condition

ν=μρ=15.10−6m2s−1=15 cSt

The dynamic viscosity of gases increases with temperature but this effect is small, of the order of
0.1% per degree; the influence of pressure is small too. One generally considers that the viscosity
of gases is constant.

Read full chapter

Further Considerations in Field Modeling


In Computational Fluid Dynamics in Fire Engineering, 2009

Dynamic Viscosity

The dynamic viscosity of the pyrolysis gas can be calculated by the proposal of Fredlund (1988),
using the proportions of pyrolysis products according to the experimental data. For the volatile
gases, the dynamic viscosity can be expressed as a function of temperature in the form of

(4.11.47)μg=μg,o+μg,mTs

where μg,o = 8.5 × 10−6 kg m−1 s−1 and μg,m = 2.95 × 10−8 kg m−1 s−1 K−1. The dynamic viscosity of
the vapor, also by Fredlund (1988), is given by

(4.11.48)μv=μv,ρ+μv,mTs
where μv,o = 8.5 × 10−6 kg m−1 s−1 and μv,m = 3.75 × 10−8 kg m−1 s−1 K−1. The dynamic viscosity of
the dry air is assumed to be constant, μi = 3.178 × 10−5 kg m−1 s−1, which is the median value for
the temperature ranging between 300°C and 1000°C (Rogers and Mayhew, 1980).

Read full chapter

Properties of Petroleum Fluids


Boyun Guo PhD, ... Xuehao Tan PhD, in Petroleum Production Engineering (Second Edition),
2017

2.3.3 Viscosity of Gas

Dynamic viscosity (μg) in centipoises (cp) is usually used in petroleum engineering. Kinematic
viscosity (vg) is related to the dynamic viscosity through density (ρg),

(2.33)vg=μgρg

Kinematic viscosity is not typically used in natural gas engineering.

Direct measurements of gas viscosity are preferred for a new gas. If gas composition and
viscosities of gas components are known, the mixing rule can be used to determine the viscosity
of the gas mixture:

(2.34)μg=∑(μgiyiMWi)∑(yiMWi)

Viscosity of gas is very often estimated with charts or correlations developed based on the charts.
Gas viscosity correlation of Carr et al., 1954 involves a two-step procedure: The gas viscosity at
temperature and atmospheric pressure is estimated first from gas-specific gravity and inorganic
compound content. The atmospheric value is then adjusted to pressure conditions by means of a
correction factor on the basis of reduced temperature and pressure state of the gas. The
atmospheric pressure viscosity (μ1) can be expressed as

(2.35)μ1=μ1HC+μ1N2+μ1CO2+μ1H2S,

where

(2.36)μ1HC=8.188×10−3−6.15×10−3log(γg)+(1.709×10−5−2.062×10−6γg)T,
(2.37)μ1N2=[9.59×10−3+8.48×10−3log(γg)]yN2,
(2.38)μ1CO2=[6.24×10−3+9.08×10−3log(γg)]yCO2,
(2.39)μ1H2S=[3.73×10−3+8.49×10−3log(γg)]yH2S,

Dempsey (1965) developed the following relation:

(2.40)μr=1n(μgμ1Tpr)=a0+a1ppr+a2ppr2+a3ppr3+Tpr(a4+a5ppr+a6ppr2+a7ppr3)+Tpr2(a8+a9p
pr+a10ppr2+a11ppr3)+Tpr3(a12+a13ppr+a14+a15ppr3),
where a0=−2.46211820; a1=2.97054714; a2=−0.28626405; a3=0.00805420; a4=2.80860949;
a5=−3.49803305; a6=0.36037302; a7=−0.01044324; a8=−0.79338568; a9=1.39643306;
a10=−0.14914493; a11=0.00441016; a12=0.08393872; a13=−0.18640885; a14=0.02033679;
a15=−0.00060958

Thus, once the value of μr is determined from the right-hand side of this equation, gas viscosity
at elevated pressure can be readily calculated using the following relation:

(2.41)μg=μ1Tpreμr

Other correlations for gas viscosity include that of Dean and Stiel (1958) and Lee et al. (1966).

Example Problem 2.3 A 0.65-specific gravity natural gas contains 10% nitrogen, 8% carbon
dioxide, and 2% hydrogen sulfide. Estimate viscosity of the gas at 10,000 psia and 180 °F.

Solution Example Problem 2.3 is solved with the spreadsheet Carr-Kobayashi-Burrows-


GasViscosity.xls, which is attached to this book. The result is shown in Table 2.3.

Table 2.3. Results Given by the Spreadsheet Carr-Kobayashi-Burrows-GasViscosity.xls

Carr-Kobayashi-Burrows-GasViscosity.xls
Description: This spreadsheet calculates gas viscosity with correlation of Carr et al.
Instruction: (1) Select a unit system; (2) update data in the Input data section; (3) review result in
the Solution section.
Input data U.S. Field units SI units
Pressure: 10,000 psia
Temperature: 180 °F
Gas-specific gravity: 0.65 air=1
Mole fraction of N2: 0.1
Mole fraction of CO2: 0.08
Mole fraction of H2S: 0.02
Solution
Pseudo-critical pressure =697.164 psia
Pseudo-critical temperature =345.357 °R
Uncorrected gas viscosity at 14.7 psia =0.012174 cp
N2 correction for gas viscosity at 14.7 psia =0.000800 cp
CO2 correction for gas viscosity at 14.7 psia =0.000363 cp
H2S correction for gas viscosity at 14.7 psia =0.000043 cp
Corrected gas viscosity at 14.7 psia (μ1) =0.013380 cp
Pseudo-reduced pressure =14.34
Pseudo-reduced temperature =1.85
In(μg/μ1 * Tpr) =1.602274
Gas viscosity =0.035843 cp
Read full chapter

Properties of Oil and Natural Gas


Boyun Guo Ph.D., ... Ali Ghalambor Ph.D., in Petroleum Production Engineering, 2007

2.3.3 Viscosity of Gas

Dynamic viscosity (μg) in centipoises (cp) is usually used in petroleum engineering. Kinematic
viscosity (vg) is related to the dynamic viscosity through density (ρg),

(2.33)vg=μgρg.

Kinematic viscosity is not typically used in natural gas engineering.

Direct measurements of gas viscosity are preferred for a new gas. If gas composition and
viscosities of gas components are known, the mixing rule can be used to determine the viscosity
of the gas mixture:

(2.34)μg=Σ(μgiyiMWi)Σ(yiMWi)

Viscosity of gas is very often estimated with charts or correlations developed based on the charts.
Gas viscosity correlation of Carr et al. 1954 involves a two-step procedure: The gas viscosity at
temperature and atmospheric pressure is estimated first from gas-specific gravity and inorganic
compound content. The atmospheric value is then adjusted to pressure conditions by means of a
correction factor on the basis of reduced temperature and pressure state of the gas. The
atmospheric pressure viscosity (μ1) can be expressed as

(2.35)μ1=μ1HC+μ1N2+μ1CO2+μ1H2S,

where

(2.36)μ1HC=8.188×10−3−6.15×10−3log(γg)+(1.709×10−5−2.062×10−6γg)T,
(2.37)μ1N2=[9.59×10−3+8.48×10−3log(γg)]yN2,
(2.38)μ1CO2=[6.24×10−3+9.08×10−3log(γg)]yCO2,
(2.39)μ1H2S=[3.73×10−3+8.49×10−3log(γg)]yH2S,

Dempsey (1965) developed the following relation:

(2.40)μr=ln(μgμ1Tpr)=a0+a1ppr+a2ppr2+a3ppr3+Tpr(a4+a5ppr)
+a6ppr2+a7ppr3)+Tpr2(a8+a9ppr+a10ppr2+a11ppr3)+Tpr3(a12+a13ppr+a14ppr2+a15ppr3),

where

a0 = −2.46211820
a1 = 2.97054714

a2 = −0.28626405

a3 = 0.00805420

a4 = 2.80860949

a5 = −3.49803305

a6 = 0.36037302

a7 = −0.01044324

a8 = −0.79338568

a9 = 1.39643306

a10 = −0.14914493

a11 = 0.00441016

a12 = 0.08393872

a13 = −0.18640885

a14 = 0.02033679

a15 = −0.00060958

Thus, once the value of μr is determined from the right-hand side of this equation, gas viscosity
at elevated pressure can be readily calculated using the following relation:

(2.41)μg=μ1Tpreμ,

Other correlations for gas viscosity include that of Dean and Stiel (1958) and Lee et al. (1966).

Example Problem 2.3

A 0.65 specific–gravity natural gas contains 10% nitrogen, 8% carbon dioxide, and 2% hydrogen
sulfide. Estimate viscosity of the gas at 10,000 psia and 180°F.

Solution Example Problem 2.3 is solved with the spreadsheet Carr-Kobayashi-Burrows-


GasViscosity.xls, which is attached to this book. The result is shown in Table 2.3.

Table 2.3. Results Given by the Spreadsheet Carr-Kobayashi-Burrows-GasViscosity.xls


Carr-Kobayashi-Burrows-GasViscosity.xls
Description: This spreadsheet calculates gas viscosity with correlation of Carr et al.
Instruction: (1) Select a unit system; (2) update data in the Input data section; (3) review
result in the Solution section.
Input data U.S. Field units SI units
Pressure: 10,000 psia
Temperature: 180 °F
Gas-specific gravity: 0.65 air = 1
Mole fraction of N2: 0.1
Mole fraction of CO2: 0.08
Mole fraction of H2S: 0.02
Solution
Pseudo-critical pressure = 697.164 psia
Pseudo-critical temperature = 345.357 °R
Uncorrected gas viscosity at 14.7 psia = 0.012174 cp
N2 correction for gas viscosity at 14.7 psia = 0.000800 cp
CO2 correction for gas viscosity at 14.7 psia = 0.000363 cp
H2S correction for gas viscosity at 14.7 psia = 0.000043 cp
Corrected gas viscosity at 14.7 psia (μ1) = 0.013380 cp
Pseudo-reduced pressure = 14.34
Pseudo-reduced temperature = 1.85
In(μg/μ1 * Tpr) = 1.602274
Gas viscosity = 0.035843 cp

Read full chapter

The Effect of Chemical Composition and Processing on


Carbon/Epoxy Laminate Quality: A Combination of Effects
F.C. Campbell, in Manufacturing Processes for Advanced Composites, 2004

7.4 Rheological Properties

Dynamic viscosity measurements were performed at 1, 2 and 5 °C min−1 on each batch of neat
resin using RDS 7700. In addition to viscosity vs. time curves, flow numbers:

flow number=∫t0tgeldt/η

were calculated from each curve. Flow is the reciprocal of viscosity when integrated as a
function of time between the starting time (t0) and the time to gellation (tgel). The test results for
the viscosity tests at a heating rate of 1 °C min−1 are summarized in Table. 6.

Table 7.6. Neat Resin Dynamic Viscosity Test Results (1 °C/min)


Batch Number 1 2(1) 3
Variation Low Flow Normal Flow High Flow
BF3 Content (wt. %) 2.2 1.1 1.1
Minimum Viscosity (poise) 10.6 6.4 1.3
Temperature at Minimum Viscosity (°C) 100 104 148
Gel Temperature at 1,000 poise (°C) 151 165 180
Flow Number (min/poise) 3.45 8.30 36.75
(1)
Control for Batches 1 and 4

The dramatic effect of the BF3 catalyst was again evident. The low flow system (batch 1)
exhibited a flow number of approximately one-half that of the normal-flow material (batch 2)
and an order of magnitude less than that of the high-flow system (batch 4).

The rest of the viscosity data further emphasize the effect of the BF3 catalyst. For example, the
minimum viscosity for the high-flow material (batch 4) was much lower than that of the low-
flow material (batch 1). The gel temperature also showed the effect of varying the catalyst
content. The low-flow system (batch 1) gelled at a lower temperature than the control (batch 2),
while the high-flow system (batch 4) gelled at a much higher temperature. The temperature at the
minimum viscosity was also significantly higher. Examining the viscosity curves shown in Fig. 4
show this difference in flow behavior. These results can be explained by the fact that the BF 3
catalyst greatly accelerates the reaction at relatively low temperatures, resulting in higher
reaction rates, lower total flow, higher minimum viscosities and lower gellation temperatures.
When the catalyst is absent, higher temperatures are necessary to initiate the chemical reactions,
thereby providing more time for total flow and higher gellation temperatures.

Sign in to download full-size image


Fig. 4. RDS Dynamic Viscosity Comparison

Read full chapter

Rheology parameters of alkali-activated geopolymeric


concrete binders
C. Leonelli, M. Romagnoli, in Handbook of Alkali-Activated Cements, Mortars and Concretes,
2015

6.2.2 Viscosity

Dynamic viscosity, denoted by the Greek letter η, is certainly the most well known of the
rheological variables. It measures the resistance with which a material deforms irreversibly when
deformed by an external force. This behavior is due to internal frictions in the material that are
generated during the deformation. The greater the viscosity of a material, the lower its
deformation per time unit under equal applied force. In rheology, a classic definition of viscosity
is given by using the Two-Plates-Model (Figure 6.10). According to this description, the sample
is placed between two surfaces ideally perfectly parallel and placed at distance h from each
other. It is also assumed that the sample has a perfect adhesion onto the two surfaces. Then a
force F→ is applied in parallel to the upper surface in order to avoid to exerting a pressure on the
material.

Sign in to download full-size image

Figure 6.10. Two-Plates-Model of viscosity.

Shear stress σ (or in some cases τ) is defined as the force F applied per unit area A:

6.1)σ=FAPa

It is expressed in Pascals (Pa) in the International System of units, similarly to pressure, although
it is not really a pressure as commonly understood. In this condition we can assess the movement
of the upper surface which, in an initial phase, accelerates until reaching a constant speed vmax
when the internal friction forces inside the sample balance the external applied force. If under
these conditions the system enters a regime of laminar motion, i.e. when the motion of the fluid
occurs with sliding of infinitesimal layers on each other without any kind of mixing of fluid even
on a microscopic scale, then it is possible to defined the shear rate γ·S−1 and the strain γ
(dimensionless). In the condition in which the speed of each sheet of fluid decreases linearly
from a maximum, in contact with the upper plate, to zero, in contact with the lower, the velocity
gradient is calculated as the ratio between the speed vmax reached by the surface in motion (the
upper) and the distance h between the two plates.

(6.2)γ·=vmaxhS−1

The shear rate is related to the flow velocity of the material: high values of γ· correspond to a
flow with high speed and vice versa. These two quantities are related to each other through a
third one known as dynamic viscosity. It is expressed as the ratio between the shear stress σ (Pa)
and the shear rate γ·S−1.

(6.3)η=σγ·Pas

The unit of measure of dynamic viscosity in the SI system is the Pascal × seconds (Pa∙s). The
most commonly used submultiple is the milliPascal × second (mPa∙s). Still surviving in common
use is the historical unit centiPoise (cP) in the CGS system, which is equivalent to mPa∙s. As a
multiple there is the Poise (P). In good practice, these units should be abandoned (Barnes et al.,
2000; Mezger, 2006).

The strain γ is defined as the ratio between the deformation of the body ξ as a result of shear
stress and the distance h between the two plates.

(6.4)γ=ξh

The viscosity is not necessarily a constant property for a material, and in fact it very rarely is
since in most cases viscosity changes under the effect of different parameters. An easy and
common classification of materials sees materials divided into ‘Newtonian’ and ‘non-
Newtonian’ fluids. The first group collects materials whose viscosity does not change with the
change of shear rate, assuming constant all the other variables that could affect the material. This
clarification is necessary because phenomena such as temperature variation, sedimentation of
any solid phases present, evaporation of the liquid phase, reactions within the material, etc., can
change the state of the system. The non-Newtonian fluids have a change in viscosity due only to
a change of the shear rate, keeping constant all the other variables. Systems with non-Newtonian
behavior are further divided as follows:

shear thinning: their viscosity decreases with increasing shear rate

shear thickening: their viscosity increases with increasing shear rate.


Since viscosity is not constant for these materials it is defined as ‘apparent viscosity’. Typically
the systems consisting of solid powder dispersed in a liquid phase in a medium–high
concentration show a non-Newtonian behavior. The variation of viscosity with shear rate is
attributed to a reorganization of the dispersed phase within the continuous phase, which
restructures favoring or hindering the flow. In the presence of superficially charged solid
particles, the phenomenon of variation of viscosity is due to rupture of the electrostatic
interactions between particles.

In Figure 6.11(a) the rheological behavior is presented as apparent viscosity vs the shear rate of
two aqueous suspensions of a metakaolin geopolymer. The two samples show a different solid
content while all the other variables of the composition remain the same. Both show a non-
Newtonian behavior, less pronounced in the system with a lower concentration. The effect of the
higher solid content produces a higher apparent viscosity which increases with time as a
consequence of the consolidation reaction. These results are also confirmed by other studies
(Criado et al., 2009), which have shown that the behavior of geopolymeric suspensions often, but
not always, follows the Bingham model (Bingham, 1916) reported below:

Sign in to download full-size image

Figure 6.11. (a) Viscosity vs shear rate of metakaolin geopolymers at two different solid content:
at lower concentration (■) and higher concentration (❍). (b) Shear stress vs.shear rate in a
geopolymer suspension with yield stress (Romagnoli et al., 2012).
(6.5)σ=σ0+Kγ·

where σ0 is the yield stress, K is a coefficient known as ‘plastic viscosity’ and γ· is the velocity
gradient.

The effect of viscosity increase due to the increase of the solid content is well known in the field
of suspensions similar to the geopolymeric ones. A model often used to describe this behavior is
the Krieger-Dougherty model (Krieger and Dougherty, 1959):

(6.6)η=ηs1−ΦΦm−ηΦm

where Φ is the volume fraction of solid contained in the suspension; Φm is the maximum volume
fraction of solid the suspension can contain corresponding to the content at which the viscosity
becomes infinite (this condition is usually ascribed to the attainment of a close-packed structure);
ηs is the viscosity of the liquid phase; and |η| the intrinsic viscosity.

he viscosity of a fluid is a measure of its resistance to deformation at a given rate. For liquids, it
corresponds to the informal concept of "thickness": for example, syrup has a higher viscosity
than water.[1]

Viscosity can be conceptualized as quantifying the frictional force that arises between adjacent
layers of fluid that are in relative motion. For instance, when a fluid is forced through a tube, it
flows more quickly near the tube's axis than near its walls. In such a case, experiments show that
some stress (such as a pressure difference between the two ends of the tube) is needed to sustain
the flow through the tube. This is because a force is required to overcome the friction between
the layers of the fluid which are in relative motion: the strength of this force is proportional to the
viscosity.

A fluid that has no resistance to shear stress is known as an ideal or inviscid fluid. Zero viscosity
is observed only at very low temperatures in superfluids. Otherwise, the second law of
thermodynamics requires all fluids to have positive viscosity;[2][3] such fluids are technically said
to be viscous or viscid. A fluid with a relatively high viscosity, such as pitch, may appear to be a
solid.

Contents
 1 Etymology
 2 Definition
o 2.1 Simple definition
o 2.2 General definition
o 2.3 Dynamic and kinematic viscosity
 3 Momentum transport
 4 Newtonian and non-Newtonian fluids
 5 In solids
 6 Measurement
 7 Units
 8 Molecular origins
o 8.1 Gases
o 8.2 Liquids
o 8.3 Mixtures, blends, and suspensions
o 8.4 Amorphous materials
o 8.5 Eddy viscosity
 9 Selected substances
o 9.1 Water
o 9.2 Air
o 9.3 Other common substances
 10 See also
 11 References
 12 Further reading
o 12.1 Undergraduate-level texts
o 12.2 Graduate-level texts
 13 External links

Etymology
The word "viscosity" is derived from the Latin "viscum", meaning mistletoe and also a viscous
glue made from mistletoe berries.[4]

Definition
Simple definition

Illustration of a planar Couette flow. Since the shearing flow is opposed by friction between
adjacent layers of fluid (which are in relative motion), a force is required to sustain the motion of
the upper plate. The relative strength of this force is a measure of the fluid's viscosity.
In a general parallel flow, the shear stress is proportional to the gradient of the velocity.

In materials science and engineering, one is often interested in understanding the forces, or
stresses, involved in the deformation of a material. For instance, if the material were a simple
spring, the answer would be given by Hooke's law, which says that the force experienced by a
spring is proportional to the distance displaced from equilibrium. Stresses which can be
attributed to the deformation of a material from some rest state are called elastic stresses. In other
materials, stresses are present which can be attributed to the rate of change of the deformation
over time. These are called viscous stresses. For instance, in a fluid such as water the stresses
which arise from shearing the fluid do not depend on the distance the fluid has been sheared;
rather, they depend on how quickly the shearing occurs.

Viscosity is the material property which relates the viscous stresses in a material to the rate of
change of a deformation (the strain rate). Although it applies to general flows, it is easy to
visualize and define in a simple shearing flow, such as a planar Couette flow.

In the Couette flow, a fluid is trapped between two infinitely large plates, one fixed and one in

parallel motion at constant speed (see illustration to the right). If the speed of the top
plate is low enough (to avoid turbulence), then in steady state the fluid particles move parallel to

it, and their speed varies from at the bottom to at the top.[5] Each layer of fluid
moves faster than the one just below it, and friction between them gives rise to a force resisting
their relative motion. In particular, the fluid applies on the top plate a force in the direction
opposite to its motion, and an equal but opposite force on the bottom plate. An external force is
therefore required in order to keep the top plate moving at constant speed.
In many fluids, the flow velocity is observed to vary linearly from zero at the bottom to

at the top. Moreover, the magnitude of the force acting on the top plate is found to be

proportional to the speed and the area of each plate, and inversely proportional

to their separation :

The proportionality factor is the viscosity of the fluid, with units of (pascal-

second). The ratio is called the rate of shear deformation or shear velocity, and is the
derivative of the fluid speed in the direction perpendicular to the plates (see illustrations to the

right). If the velocity does not vary linearly with , then the appropriate generalization is

where , and is the local shear velocity. This expression is referred to as

Newton's law of viscosity. In shearing flows with planar symmetry, it is what defines . It
is a special case of the general definition of viscosity (see below), which can be expressed in
coordinate-free form.

Use of the Greek letter mu ( ) for the viscosity is common among mechanical and

chemical engineers, as well as physicists.[6][7][8] However, the Greek letter eta ( ) is also

used by chemists, physicists, and the IUPAC.[9] The viscosity is sometimes also referred
to as the shear viscosity. However, at least one author discourages the use of this terminology,

noting that can appear in nonshearing flows in addition to shearing flows.[10]

General definition

See also: Viscous stress tensor

In very general terms, the viscous stresses in a fluid are defined as those resulting from the
relative velocity of different fluid particles. As such, the viscous stresses must depend on spatial
gradients of the flow velocity. If the velocity gradients are small, then to a first approximation
the viscous stresses depend only on the first derivatives of the velocity.[11] (For Newtonian fluids,
this is also a linear dependence.) In Cartesian coordinates, the general relationship can then be
written as

where is a viscosity tensor that maps the strain rate tensor onto the viscous

stress tensor .[12] Since the indices in this expression can vary from 1 to 3, there are 81

"viscosity coefficients" in total. However, due to spatial symmetries these coefficients


are not all independent. For instance, for isotropic Newtonian fluids, the 81 coefficients can be
reduced to 2 independent parameters. The most usual decomposition yields the standard (scalar)

viscosity and the bulk viscosity :

where is the unit tensor, and the dagger denotes the transpose.[13][14] This
equation can be thought of as a generalized form of Newton's law of viscosity.

The bulk viscosity (also called volume viscosity) expresses a type of internal friction that resists

the shearless compression or expansion of a fluid. Knowledge of is frequently not

necessary in fluid dynamics problems. For example, an incompressible fluid satisfies


and so the term containing drops out. Moreover, is often assumed to be

negligible for gases since it is in a monoatomic ideal gas.[13] One situation in which

can be important is the calculation of energy loss in sound and shock waves, described
by Stokes' law of sound attenuation, since these phenomena involve rapid expansions and
compressions.

It is worth emphasizing that the above expressions are not fundamental laws of nature, but rather
definitions of viscosity. As such, their utility for any given material, as well as means for
measuring or calculating the viscosity, must be established using separate means.

Dynamic and kinematic viscosity

In fluid dynamics, it is common to work in terms of the kinematic viscosity (also called
"momentum diffusivity"), defined as the ratio of the viscosity μ to the density of the fluid ρ. It is

usually denoted by the Greek letter nu (ν) and has units :

Consistent with this nomenclature, the viscosity is frequently called the dynamic
viscosity or absolute viscosity, and has units force × time/area.

Momentum transport
See also: Transport phenomena
Transport theory provides an alternate interpretation of viscosity in terms of momentum
transport: viscosity is the material property which characterizes momentum transport within a
fluid, just as thermal conductivity characterizes heat transport, and (mass) diffusivity

characterizes mass transport.[15] To see this, note that in Newton's law of viscosity, , the

shear stress has units equivalent to a momentum flux, i.e. momentum per unit time per

unit area. Thus, can be interpreted as specifying the flow of momentum in the
direction from one fluid layer to the next. Per Newton's law of viscosity, this momentum flow
occurs across a velocity gradient, and the magnitude of the corresponding momentum flux is
determined by the viscosity.

The analogy with heat and mass transfer can be made explicit. Just as heat flows from high
temperature to low temperature and mass flows from high density to low density, momentum
flows from high velocity to low velocity. These behaviors are all described by compact
expressions, called constitutive relations, whose one-dimensional forms are given here:

where is the density, and are the mass and heat fluxes, and

and are the mass diffusivity and thermal conductivity.[16]

The fact that mass, momentum, and energy (heat) transport are among the most relevant
processes in continuum mechanics is not a coincidence: these are among the few physical
quantities that are conserved at the microscopic level in interparticle collisions. Thus, rather than
being dictated by the fast and complex microscopic interaction timescale, their dynamics occurs
on macroscopic timescales, as described by the various equations of transport theory and
hydrodynamics.

Newtonian and non-Newtonian fluids

Viscosity, the slope of each line, varies among materials.


Newton's law of viscosity is not a fundamental law of nature, but rather a constitutive equation

(like Hooke's law, Fick's law, and Ohm's law) which serves to define the viscosity . Its
form is motivated by experiments which show that for a wide range of fluids, is
independent of strain rate. Such fluids are called Newtonian. Gases, water, and many common
liquids can be considered Newtonian in ordinary conditions and contexts. However, there are
many non-Newtonian fluids that significantly deviate from this behavior. For example:

 Shear-thickening liquids, whose viscosity increases with the rate of shear strain.
 Shear-thinning liquids, whose viscosity decreases with the rate of shear strain.
 Thixotropic liquids, that become less viscous over time when shaken, agitated, or
otherwise stressed.
 Rheopectic (dilatant) liquids, that become more viscous over time when shaken, agitated,
or otherwise stressed.
 Bingham plastics that behave as a solid at low stresses but flow as a viscous fluid at high
stresses.

The Trouton ratio or Trouton's ratio is the ratio of extensional viscosity to shear viscosity.[17] For
a Newtonian fluid, the Trouton ratio is 3.[18]

Shear-thinning liquids are very commonly, but misleadingly, described as thixotropic.[19]

Even for a Newtonian fluid, the viscosity usually depends on its composition and temperature.
For gases and other compressible fluids, it depends on temperature and varies very slowly with
pressure. The viscosity of some fluids may depend on other factors. A magnetorheological fluid,
for example, becomes thicker when subjected to a magnetic field, possibly to the point of
behaving like a solid.

In solids
The viscous forces that arise during fluid flow must not be confused with the elastic forces that
arise in a solid in response to shear, compression or extension stresses. While in the latter the
stress is proportional to the amount of shear deformation, in a fluid it is proportional to the rate
of deformation over time. (For this reason, Maxwell used the term fugitive elasticity for fluid
viscosity.)

However, many liquids (including water) will briefly react like elastic solids when subjected to
sudden stress. Conversely, many "solids" (even granite) will flow like liquids, albeit very slowly,
even under arbitrarily small stress.[20] Such materials are therefore best described as possessing
both elasticity (reaction to deformation) and viscosity (reaction to rate of deformation); that is,
being viscoelastic.

Indeed, some authors have claimed that amorphous solids, such as glass and many polymers, are
actually liquids with a very high viscosity (greater than 1012 Pa·s). [21] However, other authors
dispute this hypothesis, claiming instead that there is some threshold for the stress, below which
most solids will not flow at all,[22] and that alleged instances of glass flow in window panes of old
buildings are due to the crude manufacturing process of older eras rather than to the viscosity of
glass.[23]

Viscoelastic solids may exhibit both shear viscosity and bulk viscosity. The extensional viscosity
is a linear combination of the shear and bulk viscosities that describes the reaction of a solid
elastic material to elongation. It is widely used for characterizing polymers.

In geology, earth materials that exhibit viscous deformation at least three orders of magnitude
greater than their elastic deformation are sometimes called rheids.[24]

Measurement
Main article: Viscometer

Viscosity is measured with various types of viscometers and rheometers. A rheometer is used for
those fluids that cannot be defined by a single value of viscosity and therefore require more
parameters to be set and measured than is the case for a viscometer. Close temperature control of
the fluid is essential to acquire accurate measurements, particularly in materials like lubricants,
whose viscosity can double with a change of only 5 °C.

For some fluids, the viscosity is constant over a wide range of shear rates (Newtonian fluids).
The fluids without a constant viscosity (non-Newtonian fluids) cannot be described by a single
number. Non-Newtonian fluids exhibit a variety of different correlations between shear stress
and shear rate.

One of the most common instruments for measuring kinematic viscosity is the glass capillary
viscometer.

In coating industries, viscosity may be measured with a cup in which the efflux time is measured.
There are several sorts of cup – such as the Zahn cup and the Ford viscosity cup – with the usage
of each type varying mainly according to the industry. The efflux time can also be converted to
kinematic viscosities (centistokes, cSt) through the conversion equations.[25]

Also used in coatings, a Stormer viscometer uses load-based rotation in order to determine
viscosity. The viscosity is reported in Krebs units (KU), which are unique to Stormer
viscometers.

Vibrating viscometers can also be used to measure viscosity. Resonant, or vibrational


viscometers work by creating shear waves within the liquid. In this method, the sensor is
submerged in the fluid and is made to resonate at a specific frequency. As the surface of the
sensor shears through the liquid, energy is lost due to its viscosity. This dissipated energy is then
measured and converted into a viscosity reading. A higher viscosity causes a greater loss of
energy.[citation needed]

Extensional viscosity can be measured with various rheometers that apply extensional stress.
Volume viscosity can be measured with an acoustic rheometer.

Apparent viscosity is a calculation derived from tests performed on drilling fluid used in oil or
gas well development. These calculations and tests help engineers develop and maintain the
properties of the drilling fluid to the specifications required.

Units
The SI unit of dynamic viscosity is the pascal-second (Pa·s), or equivalently kilogram per meter
per second (kg·m−1·s−1). The CGS unit is called the poise[26] (P), named after Jean Léonard Marie
Poiseuille. It is commonly expressed, particularly in ASTM standards, as centipoise (cP) since
the latter is equal to the SI multiple millipascal seconds (mPa·s).

The SI unit of kinematic viscosity is square meter per second (m2/s), whereas the CGS unit for
kinematic viscosity is the stokes (St), named after Sir George Gabriel Stokes.[27] In U.S. usage,
stoke is sometimes used as the singular form. The submultiple centistokes (cSt) is often used
instead.

The reciprocal of viscosity is fluidity, usually symbolized by or , depending on


−1 −1
the convention used, measured in reciprocal poise (P , or cm·s·g ), sometimes called the rhe.
Fluidity is seldom used in engineering practice.

Nonstandard units include the reyn, a British unit of dynamic viscosity.[citation needed] In the
automotive industry the viscosity index is used to describe the change of viscosity with
temperature.

At one time the petroleum industry relied on measuring kinematic viscosity by means of the
Saybolt viscometer, and expressing kinematic viscosity in units of Saybolt universal seconds
(SUS).[28] Other abbreviations such as SSU (Saybolt seconds universal) or SUV (Saybolt
universal viscosity) are sometimes used. Kinematic viscosity in centistokes can be converted
from SUS according to the arithmetic and the reference table provided in ASTM D 2161.[29]

Molecular origins
In general, the viscosity of a system depends in detail on how the molecules constituting the
system interact. There are no simple but correct expressions for the viscosity of a fluid. The
simplest exact expressions are the Green–Kubo relations for the linear shear viscosity or the
transient time correlation function expressions derived by Evans and Morriss in 1985.[30]
Although these expressions are each exact, calculating the viscosity of a dense fluid using these
relations currently requires the use of molecular dynamics computer simulations. On the other
hand, much more progress can be made for a dilute gas. Even elementary assumptions about how
gas molecules move and interact lead to a basic understanding of the molecular origins of
viscosity. More sophisticated treatments can be constructed by systematically coarse-graining the
equations of motion of the gas molecules. An example of such a treatment is Chapman–Enskog
theory, which derives expressions for the viscosity of a dilute gas from the Boltzmann equation.
[31]

Momentum transport in gases is generally mediated by discrete molecular collisions, and in


liquids by attractive forces which bind molecules close together.[15] Because of this, the dynamic
viscosities of liquids are typically much larger than those of gases.

Amorphous materials

Common glass viscosity curves[54]

In the high and low temperature limits, viscous flow in amorphous materials (e.g. in glasses and
melts)[55][56][57] has the Arrhenius form:

where Q is a relevant activation energy, given in terms of molecular parameters; T is


temperature; R is the molar gas constant; and A is approximately a constant. The activation
energy Q takes a different value depending on whether the high or low temperature limit is being
considered: it changes from a high value QH at low temperatures (in the glassy state) to a low
value QL at high temperatures (in the liquid state).

Common logarithm of viscosity against temperature for B2O3, showing two regimes
For intermediate temperatures, varies nontrivially with temperature and the simple
Arrhenius form fails. On the other hand, the two-exponential equation

where , , , are all constants, provides a good fit to


experimental data over the entire range of temperatures, while at the same time reducing to the
correct Arrhenius form in the low and high temperature limits. Besides being a convenient fit to
data, the expression can also be derived from various theoretical models of amorphous materials
at the atomic level.[57]

Eddy viscosity

In the study of turbulence in fluids, a common practical strategy is to ignore the small-scale
vortices (or eddies) in the motion and to calculate a large-scale motion with an effective
viscosity, called the "eddy viscosity", which characterizes the transport and dissipation of energy
in the smaller-scale flow (see large eddy simulation).[58][59] In contrast to the viscosity of the fluid
itself, which must be positive by the second law of thermodynamics, the eddy viscosity can be
negative.[60][61]

Selected substances
Main article: List of viscosities

In the University of Queensland pitch drop experiment, pitch has been dripping slowly through a
funnel since 1927, at a rate of one drop roughly every decade. In this way the viscosity of pitch
has been determined to be approximately 230 billion (2.3×1011) times that of water.[62]
Observed values of viscosity vary over several orders of magnitude, even for common
substances. For instance, a 70% sucrose (sugar) solution has a viscosity over 400 times that of
water, and 26000 times that of air.[63] More dramatically, pitch has been estimated to have a
viscosity 230 billion times that of water.[62]

Water

The dynamic viscosity of water is about 0.89 mPa·s at room temperature (25 °C). As a
function of temperature, the dynamic viscosity can be estimated using the semi-empirical
relation:

where A = 2.414×10−5 Pa·s, B = 247.8 K, and C = 140 K.[citation needed]

Experimentally determined values of the dynamic viscosity at various temperatures are given
below.

Viscosity of water
at various temperatures[63]
Temperature (°C) Viscosity (mPa·s)
10 1.3059
20 1.0016
30 0.79722
50 0.54652
70 0.40355
90 0.31417

Air

Under standard atmospheric conditions (25 °C and pressure of 1 bar), the dynamic viscosity of
air is 18.5 μPa·s, roughly 50 times smaller than the viscosity of water at the same temperature.
Except at very high pressure, the viscosity of air depends mostly on the temperature.

Other common substances


Honey being drizzled
Substance Viscosity (mPa·s) Temperature (°C)
Benzene 0.604 25
[63]
Water 1.0016 20
Mercury 1.526 25
[64]
Whole milk 2.12 20
Olive oil[64] 56.2 26

Honey[65] 2000- 20
10000

Ketchup[a][66] 5000- 25
20000
Peanut butter[a][67]
104-106
Pitch[62] 2.3×1011 10-30 (variable)

1.

These materials are highly non-Newtonian.

You might also like