You are on page 1of 113

Contact of these minute metal projections should be minimized; however, some contact always

occurs and results in normal wear of the metal surfaces. If contact is not minimized, heat is
generated when the metal parts touch. The heat causes local welding and transfer of metal, which
creates scuffing or seizing of the equipment. These actions are called adhesive wear. The oil property
that governs the thickness of the separating oil film is the viscosity. Viscosity The commonly used
kinematic viscosity is defined as a measure of the restrictive flow of a fluid under gravitational force.
The “cgs” unit of kinematic viscosity (one centimeter squared per second), is called one stoke (St).
The SI unit for kinematic viscosity is one meter squared per second and is equivalent to 10,000 St.
Usually, centistokes (cSt) is used (1 cSt = 0.01 St = 1 mm2 /s). The absolute or dynamic viscosity is
equal to the kinematic viscosity, multiplied by the density of the fluid. It is usually expressed in
centipoise (cP) (1 cP = 0.001 Pa.s). Viscosity index (VI) The viscosity of lubricating oil changes with
temperature and the rate of change depends on the composition of the oil. Naphthenic base oils
change more than paraffinic base oils. Certain synthetic lubricants change much less than paraffinic
oils. To assess this lubricating oil property, the American Society for Testing and Materials (ASTM)
created a method to provide a number called the Viscosity Index (VI). The VI correlates the amount
of viscosity change for a given oil, compared to two reference oils having the highest and lowest
viscosity indices at the time the VI scale was first introduced (1929). A standard paraffinic oil was
given a VI of 100 and a standard naphthenic oil a VI of 0. Figure 1 shows the relationship between
viscosities at 40°C and 100°C. The method has been updated and revised several times to include VI
values higher than 100. The first and most important task of lubricating oil is to keep moving metal
parts separated from each other, thus avoiding metal-to-metal contact, which leads to destructive
wear. Even finely machined metal surfaces have a certain roughness. Figure 1: Schematic
Representation of Viscosity Index VI of unknown oil VI = 0 VI = 100 Unknown oil Viscosity 40°C 100°C
100 VI scale 0 viscosity classifications marine lubricants information bulletin 6 A low VI means a
relatively large viscosity change with temperature and a high VI denotes a smaller change of viscosity
with temperature. Hence, the VI of an oil is important in applications where an appreciable change
in temperature of the lubricating oil could affect the startup or operating characteristics of the
equipment. Deck machinery and emergency equipment are examples of typical applications onboard
ships. Viscosity classification As the selection of the proper viscosity grade is extremely important,
various viscosity classification systems have been developed over the years. The viscosity
classification for engine oils was developed by the Society for Automotive Engineers (SAE) in 1911.
After many revisions and updates, this classification system is still in place. The current SAE J300
Viscosity Classification is shown on the next page in Table 1. SAE grades 0W through 25W, where W
stands for winter, have a maximum viscosity specified at low temperatures (–5°C through –35°C), to
ensure easy starting under low temperature conditions, and a minimum viscosity requirement at
100°C to ensure satisfactory lubrication at the final operating temperature. Only SAE grades 20W
through 60W have limits set at 100°C, because these grades are not intended to be used in low
temperature conditions. For marine applications, monograde oils (i.e., oils without the addition of VI
improvers of SAE 30 or SAE 40) are used because of the steady operating conditions in a ship’s
engine room. Conversely, automotive oils are normally formulated by adding VI improvers to
provide multigrade performance and thus deliver excellent temperature/viscosity relations. VI
improvers are very large molecules, which are chemically made by linking together smaller
molecules into so-called polymers. The use of these special polymers makes it possible to meet the
low temperature viscosity requirements of the W grades, as well as the high temperature
requirements of the non-W grades. In a 15W-40 multigrade engine oil, the typical viscosities are:
This example shows that the high VI offers a relatively small change in viscosity with temperature,
and, as a result of the high VI, the multigrade oil meets the 15W grade low temperature viscosity
requirements, as well as the 40 grade high temperature viscosity requirements. Viscosity
classification: industrial oils Many different viscosity classification systems have been used in the
past in different parts of the world. It has been difficult to reach agreement on the number of
different grades to be included, the viscosity limits for these grades, and the temperature at which
the viscosity should be specified. It is only since 1972 that a worldwide viscosity classification system
for industrial lubricants came into place. The current ISO 3448 viscosity classification system, which
is also adopted by the ASTM, is shown on the next page in Table 2. The classification is based on a
series of viscosity grades, each being approximately 50% more viscous than its preceding grade,
while the viscosity deviation within a grade is plus or minus 10% of the nominal viscosity of that
grade. Used oil viscosities The viscosity is determined on every oil sample tested in Chevron’s FAST™
used oil analysis service. Used lubricating oils may show an increase of the viscosity because of
oxidation/nitration or contamination such as the soot loading of a diesel engine oil. Dilution with
high viscosity heavy fuel oil or the use of a higher viscosity grade lubricant are other possible causes.
A viscosity decrease of a used lubricating oil may be related to the use of a lower viscosity grade
lubricant or dilution with low viscosity fuel oil. Viscosity results of used lubricating oil samples are
compared to the original equipment manufacturer’s (OEM’s) requirements whenever possible. If this
is not feasible, the generally accepted limit for viscosity change is ±15% of the fresh oil value at 40°C.
For large diesel engines, an increase of the engine oil’s kinematic viscosity at 40°C up to 45% is still
accepted. ■ Viscosity at –15°C, cP 3,000 Viscosity at 40°C, mm2 /s (cSt) 105 Viscosity at 100°C,
mm2 /s (cSt) 14 Viscosity Index 135 2 | Information Bulletin 6: Viscosity Classifications Information
Bulletin 6: Viscosity Classifications | 3 Table 2: Industrial lubricant viscosity classification Viscosity
system grade Mid-point viscosity, Kinematic viscosity limits, mm2 /s (cSt), at 40°C ISO standard 3448
mm2 /s (cSt), at 40°C ASTM D 2422 Min Max ISO VG 2 2.2 1.98 2.42 ISO VG 3 3.2 2.88 3.52 ISO VG 5
4.6 4.14 5.06 ISO VG 7 6.8 6.12 7.48 ISO VG 10 10 9.00 11.0 ISO VG 15 15 13.5 16.5 ISO VG 22 22 19.8
24.2 ISO VG 32 32 28.8 35.2 ISO VG 46 46 41.4 50.6 ISO VG 68 68 61.2 74.8 ISO VG 100 100 90.0 110
ISO VG 150 150 135 165 ISO VG 220 220 198 242 ISO VG 320 320 288 352 ISO VG 460 460 414 506
ISO VG 680 680 612 748 ISO VG 1000 1,000 900 1,100 ISO VG 1500 1,500 1,300 1,650 Table 1: Engine
oil viscosity classification — SAE J300 revised May 2004 SAE viscosity Low- Low-temperature Low-
shear-rate Low-shear-rate High-shear-rate grade temperature (°C) pumping kinematic kinematic
viscosity (°C) cranking viscosity3 viscosity4 viscosity5 (mPa.s) at viscosity2 , mPa.s max. with (mm2
/s) at (mm2 /s) at 150°C and mPa.s max. no yield stress 100°C min. 100°C max. 106 s–1 min. 0 W
6,200 at –35 60,000 at –40 3.8 — — 5 W 6,600 at –30 60,000 at –35 3.8 — — 10 W 7,000 at –25
60,000 at –30 4.1 — — 15 W 7,000 at –20 60,000 at –25 5.6 — — 20 W 9,500 at –15 60,000 at –20
5.6 — — 25 W 13,000 at –10 60,000 at –15 9.3 — — 20 — — 5.6 < 9.3 2.6 30 — — 9.3 < 12.5 2.9 40
— — 12.5 < 16.3 2.9 (0W-40, 5W-50, and 10W-40 grades) 40 — — 12.5 < 16.3 3.7 (15W-40, 20W-40,
25W40, 40 grades) 50 — — 16.3 < 21.9 3.7 60 — — 21.9 < 26.1 3.7 1. mPa.s = 1cP; 1 mm2 /s = 1 cSt.
All values are critical specifications as defined by ASTM D 3244. 2. ASTM D 5293 3. ASTM D 4684
Note: The presence of any yield stress detectable by this method constitutes a failure regardless of
viscosity. 4. ASTM D 445 5. ASTM D 4683, CE
Main menu











Search Go

 Create account
 Log in
Personal tools


Contents
 hide

(Top)


Etymology


Definition
Toggle Definition subsection

Momentum transport


Newtonian and non-Newtonian fluids


In solids


Measurement


Units


Molecular origins
Toggle Molecular origins subsection

Prediction


Selected substances
Toggle Selected substances subsection

See also


References
Toggle References subsection

External links

Viscosity
76 languages
 Article
 Talk
 Read
 Edit
 View history
Tools











From Wikipedia, the free encyclopedia

Viscosity

A simulation of liquids with different viscosities. The liquid on the left

has lower viscosity than the liquid on the right.

Common symbols η, μ

Derivations from μ = G·t


other quantities
Part of a series on

Continuum mechanics

Fick's laws of diffusion

Laws

Solid mechanics

Fluid mechanics

Fluids
 Statics · Dynamics
 Archimedes' principle · Bernoulli's
principle
 Navier–Stokes equations
 Poiseuille equation · Pascal's law
 Viscosity
o (Newtonian · non-Newtonian)
 Buoyancy · Mixing · Pressure
Liquids
 Adhesion
 Capillary action
 Chromatography
 Cohesion (chemistry)
 Surface tension
Gases
 Atmosphere
 Boyle's law
 Charles's law
 Combined gas law
 Fick's law
 Gay-Lussac's law
 Graham's law
Plasma

Rheology

Scientists

 v
 t
 e
The 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 quantifies the internal frictional force between adjacent layers of fluid that
are in relative motion. For instance, when a viscous fluid is forced through a tube, it
flows more quickly near the tube's axis than near its walls. Experiments show that
some stress (such as a pressure difference between the two ends of the tube) is
needed to sustain the flow. This is because a force is required to overcome the
friction between the layers of the fluid which are in relative motion. For a tube with a
constant rate of flow, the strength of the compensating force is proportional to the
fluid's viscosity.
In general, viscosity depends on a fluid's state, such as its temperature, pressure,
and rate of deformation. However, the dependence on some of these properties is
negligible in certain cases. For example, the viscosity of a Newtonian fluid does not
vary significantly with the rate of deformation. Zero viscosity (no resistance to shear
stress) is observed only at very low temperatures in superfluids; otherwise,
the second law of thermodynamics requires all fluids to have positive viscosity. [2][3] A
fluid that has zero viscosity is called ideal or inviscid.

Etymology[edit]
The word "viscosity" is derived from the Latin viscum ("mistletoe"). Viscum also
referred to a viscous glue derived from mistletoe berries.[4]

Definition[edit]
Dynamic viscosity[edit]

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 dynamic viscosity of the fluid, often simply
referred to as the viscosity. It is denoted by the Greek letter mu (μ). The dynamic
viscosity has the dimensions , therefore resulting in the SI units and the derived
units:
 pressure multiplied by time.
The aforementioned 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 normal vector of 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 dynamic viscosity (sometimes also
called the absolute viscosity) is common among mechanical and chemical
engineers, as well as mathematicians and physicists. [6][7][8] However,
the Greek letter eta () is also used by chemists, physicists, and
the IUPAC.[9] The viscosity  is sometimes also called the shear viscosity.
However, at least one author discourages the use of this terminology,
noting that  can appear in non-shearing flows in addition to shearing
flows.[10]
Kinematic viscosity[edit]
In fluid dynamics, it is sometimes more appropriate to work in terms
of kinematic viscosity (sometimes also called the momentum diffusivity),
defined as the ratio of the dynamic viscosity (μ) over the density of the
fluid (ρ). It is usually denoted by the Greek letter nu (ν):
and has the dimensions , therefore resulting in the SI units and
the derived units:
 specific energy multiplied by time.
General definition[edit]
See also: Viscous stress tensor and Volume viscosity
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 velocity
gradient 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, assuming that the
viscosity rank-4 tensor is isotropic reduces these 81
coefficients to three independent parameters , , :
and furthermore, it is assumed that no viscous forces may
arise when the fluid is undergoing simple rigid-body
rotation, thus , leaving only two independent parameters.
 The most usual decomposition is in terms of the standard
[11]

(scalar) viscosity  and the bulk viscosity  such that  and . In


vector notation this appears as:
where  is the unit tensor, and the dagger  denotes
the transpose.[10][13] 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 monatomic ideal gas.[10] 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.
The defining equations for viscosity are not fundamental
laws of nature, so their usefulness, as well as methods
for measuring or calculating the viscosity, must be
established using separate means. A potential issue is
that viscosity depends, in principle, on the full
microscopic state of the fluid, which encompasses the
positions and momenta of every particle in the system.
[14]
 Such highly detailed information is typically not
available in realistic systems. However, under certain
conditions most of this information can be shown to be
negligible. In particular, for Newtonian fluids near
equilibrium and far from boundaries (bulk state), the
viscosity depends only space- and time-dependent
macroscopic fields (such as temperature and density)
defining local equilibrium.[14][15]
Nevertheless, viscosity may still carry a non-negligible
dependence on several system properties, such as
temperature, pressure, and the amplitude and
frequency of any external forcing. Therefore, precision
measurements of viscosity are only defined with respect
to a specific fluid state.[16] To standardize comparisons
among experiments and theoretical models, viscosity
data is sometimes extrapolated to ideal limiting cases,
such as the zero shear limit, or (for gases) the zero
density limit.

Momentum transport[edit]
See also: Transport phenomena
Transport theory provides an alternative 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.[17] This
perspective is implicit in Newton's law of viscosity, ,
because 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.[18] 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[edit]
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 (dilatant) 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 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.
Trouton's ratio is the ratio of extensional
viscosity to shear viscosity. For a Newtonian fluid,
the Trouton ratio is 3.[19][20] Shear-thinning liquids are
very commonly, but misleadingly, described as
thixotropic.[21]
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[edit]
The viscous forces that arise during fluid flow are
distinct from the elastic forces that occur 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.[22] Such materials are best
described as viscoelastic—that is, possessing both
elasticity (reaction to deformation) and viscosity
(reaction to rate of deformation).
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.[23]

Measurement[edit]
Main article: Viscometer
Viscosity is measured with various types
of viscometers and rheometers. A rheometer is
used for 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 obtain accurate measurements,
particularly in materials like lubricants, whose
viscosity can double with a change of only 5 °C.[24]
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.
Also used in coatings, a Stormer
viscometer employs load-based rotation 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.
Nanoviscosity (viscosity sensed by nanoprobes) can
be measured by fluorescence correlation
spectroscopy.[25]

Units[edit]
The SI unit of dynamic viscosity is the newton-
second per square meter (N·s/m2), also frequently
expressed in the equivalent forms pascal-
second (Pa·s), kilogram per meter per second
(kg·m−1·s−1) and poiseuille (Pl). The CGS unit is
the poise (P, or g·cm−1·s−1 = 0.1 Pa·s),[26] named
after Jean Léonard Marie Poiseuille. It is commonly
expressed, particularly in ASTM standards,
as centipoise (cP). The centipoise is convenient
because the viscosity of water at 20 °C is about 1
cP, and one centipoise is equal to the SI millipascal
second (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, or cm2·s−1 =
0.0001 m2·s−1), 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, 1 cSt = 1 mm2·s−1 = 10−6 m2·s−1.
The kinematic viscosity of water at 20 °C is about 1
cSt.
The most frequently used systems of US customary,
or Imperial, units are the British Gravitational (BG)
and English Engineering (EE). In the BG system,
dynamic viscosity has units of pound-seconds per
square foot (lb·s/ft2), and in the EE system it has
units of pound-force-seconds per square foot
(lbf·s/ft2). The pound and pound-force are
equivalent; the two systems differ only in how force
and mass are defined. In the BG system the pound
is a basic unit from which the unit of mass (the slug)
is defined by Newton's Second Law, whereas in the
EE system the units of force and mass (the pound-
force and pound-mass respectively) are defined
independently through the Second Law using
the proportionality constant gc.
Kinematic viscosity has units of square feet per
second (ft2/s) in both the BG and EE systems.
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.
The reciprocal of viscosity is fluidity, usually
symbolized by  or , depending on the convention
used, measured in reciprocal poise (P−1, or cm·s·g−1),
sometimes called the rhe. Fluidity is seldom used
in engineering practice.[citation needed]
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.

Molecular origins[edit]
Momentum transport in gases is mediated by
discrete molecular collisions, and in liquids by
attractive forces that bind molecules close together.
[17]
 Because of this, the dynamic viscosities of liquids
are typically much larger than those of gases. In
addition, viscosity tends to increase with
temperature in gases and decrease with
temperature in liquids.
Above the liquid-gas critical point, the liquid and gas
phases are replaced by a single supercritical phase.
In this regime, the mechanisms of momentum
transport interpolate between liquid-like and gas-like
behavior. For example, along a
supercritical isobar (constant-pressure surface), the
kinematic viscosity decreases at low temperature
and increases at high temperature, with a minimum
in between.[29][30] A rough estimate for the value at the
minimum is
where  is the Planck constant,  is the electron
mass, and  is the molecular mass.[30]
In general, however, the viscosity of a system
depends in detail on how the molecules
constituting the system interact, and there are
no simple but correct formulas for it. 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 1988.[31] 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. Somewhat
more progress can be made for a dilute gas, as
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.[15]
Pure gases[edit]
See also: Kinetic theory of gases
Elementary calculation of viscosity for a dilute gas
Viscosity in gases arises principally from
the molecular diffusion that transports
momentum between layers of flow. An
elementary calculation for a dilute gas at
temperature  and density  gives
where  is the Boltzmann constant,  the
molecular mass, and  a numerical
constant on the order of . The quantity ,
the mean free path, measures the
average distance a molecule travels
between collisions. Even without a
priori knowledge of , this expression has
nontrivial implications. In particular,
since  is typically inversely proportional to
density and increases with
temperature,  itself should increase with
temperature and be independent of
density at fixed temperature. In fact, both
of these predictions persist in more
sophisticated treatments, and accurately
describe experimental observations. By
contrast, liquid viscosity typically
decreases with temperature.[17][32]
For rigid elastic spheres of
diameter ,  can be computed, giving
In this case  is independent of
temperature, so . For more
complicated molecular models,
however,  depends on temperature in
a non-trivial way, and simple kinetic
arguments as used here are
inadequate. More fundamentally, the
notion of a mean free path becomes
imprecise for particles that interact
over a finite range, which limits the
usefulness of the concept for
describing real-world gases.[33]
Chapman–Enskog theory[edit]
Main article: Chapman–Enskog
theory
A technique developed by Sydney
Chapman and David Enskog in the
early 1900s allows a more refined
calculation of .[15] It is based on
the Boltzmann equation, which
provides a statistical description of a
dilute gas in terms of intermolecular
interactions.[34] The technique allows
accurate calculation of  for molecular
models that are more realistic than
rigid elastic spheres, such as those
incorporating intermolecular
attractions. Doing so is necessary to
reproduce the correct temperature
dependence of , which experiments
show increases more rapidly than
the  trend predicted for rigid elastic
spheres.[17] Indeed, the Chapman–
Enskog analysis shows that the
predicted temperature dependence
can be tuned by varying the
parameters in various molecular
models. A simple example is the
Sutherland model,[a] which describes
rigid elastic spheres with weak mutual
attraction. In such a case, the
attractive force can be
treated perturbatively, which leads to
a simple expression for :
where  is independent of
temperature, being determined
only by the parameters of the
intermolecular attraction. To
connect with experiment, it is
convenient to rewrite as
where  is the viscosity at
temperature .[35] If  is known
from experiments at  and at
least one other temperature,
then  can be calculated.
Expressions for  obtained in
this way are qualitatively
accurate for a number of
simple gases. Slightly more
sophisticated models, such as
the Lennard-Jones potential,
may provide better agreement
with experiments, but only at
the cost of a more opaque
dependence on temperature.
In some systems, the
assumption of spherical
symmetry must be abandoned
as well, as is the case for
vapors with highly polar
molecules like H2O. In these
cases, the Chapman–Enskog
analysis is significantly more
complicated.[36][37]
Bulk viscosity[edit]
In the kinetic-molecular
picture, a non-zero bulk
viscosity arises in gases
whenever there are non-
negligible relaxational
timescales governing the
exchange of energy between
the translational energy of
molecules and their internal
energy,
e.g. rotational and vibrational.
As such, the bulk viscosity
is  for a monatomic ideal gas,
in which the internal energy of
molecules is negligible, but is
nonzero for a gas like carbon
dioxide, whose molecules
possess both rotational and
vibrational energy.[38][39]
Pure liquids[edit]
See also: Temperature
dependence of liquid viscosity
Video showing three liquids with
different viscosities

Experiment showing the behavior


of a viscous fluid with blue dye for
visibility

In contrast with gases, there is


no simple yet accurate picture
for the molecular origins of
viscosity in liquids.
At the simplest level of
description, the relative motion
of adjacent layers in a liquid is
opposed primarily by attractive
molecular forces acting across
the layer boundary. In this
picture, one (correctly)
expects viscosity to decrease
with increasing temperature.
This is because increasing
temperature increases the
random thermal motion of the
molecules, which makes it
easier for them to overcome
their attractive interactions.[40]
Building on this visualization,
a simple theory can be
constructed in analogy with
the discrete structure of a
solid: groups of molecules in a
liquid are visualized as
forming "cages" which
surround and enclose single
molecules.[41] These cages can
be occupied or unoccupied,
and stronger molecular
attraction corresponds to
stronger cages. Due to
random thermal motion, a
molecule "hops" between
cages at a rate which varies
inversely with the strength of
molecular attractions.
In equilibrium these "hops" are
not biased in any direction. On
the other hand, in order for
two adjacent layers to move
relative to each other, the
"hops" must be biased in the
direction of the relative
motion. The force required to
sustain this directed motion
can be estimated for a given
shear rate, leading to
 
 

   
(1)

where  is the Avogadro
constant,  is the Planck
constant,  is the volume of
a mole of liquid, and  is
the normal boiling point.
This result has the same
form as the well-known
empirical relation
 
 
(2
)
 

where  and  are
constants fit from data.
[41][42]
 On the other hand,
several authors
express caution with
respect to this model.
Errors as large as 30%
can be encountered
using equation (1),
compared with fitting
equation (2) to
experimental data.
[41]
 More fundamentally,
the physical
assumptions
underlying equation (1)
have been criticized.
[43]
 It has also been
argued that the
exponential
dependence in
equation (1) does not
necessarily describe
experimental
observations more
accurately than
simpler, non-
exponential
expressions.[44][45]
In light of these
shortcomings, the
development of a less
ad hoc model is a
matter of practical
interest. Foregoing
simplicity in favor of
precision, it is possible
to write rigorous
expressions for
viscosity starting from
the fundamental
equations of motion for
molecules. A classic
example of this
approach is Irving–
Kirkwood theory.[46] On
the other hand, such
expressions are given
as averages over
multiparticle correlation
functions and are
therefore difficult to
apply in practice.
In general, empirically
derived expressions
(based on existing
viscosity
measurements) appear
to be the only
consistently reliable
means of calculating
viscosity in liquids.[47]
Mixtures and
blends[edit]
See also: Viscosity
models for mixtures
Gaseous mixtures[edit]
The same molecular-
kinetic picture of a
single component gas
can also be applied to
a gaseous mixture. For
instance, in the
Chapman–Enskog
approach the
viscosity  of a binary
mixture of gases can
be written in terms of
the individual
component viscosities ,
their respective volume
fractions, and the
intermolecular
interactions.[15] As for
the single-component
gas, the dependence
of  on the parameters
of the intermolecular
interactions enters
through various
collisional integrals
which may not be
expressible in terms of
elementary functions.
To obtain usable
expressions for  which
reasonably match
experimental data, the
collisional integrals
typically must be
evaluated using some
combination of analytic
calculation and
empirical fitting. An
example of such a
procedure is the
Sutherland approach
for the single-
component gas,
discussed above.
Blends of liquids[edit]
As for pure liquids, the
viscosity of a blend of
liquids is difficult to
predict from molecular
principles. One method
is to extend the
molecular "cage"
theory presented
above for a pure liquid.
This can be done with
varying levels of
sophistication. One
expression resulting
from such an analysis
is the Lederer–
Roegiers equation for a
binary mixture:
where  is an
empirical
parameter,
and  and  are the
respective mole
fractions and
viscosities of the
component liquids.
[48]
Since blending is
an important
process in the
lubricating and oil
industries, a variety
of empirical and
propriety equations
exist for predicting
the viscosity of a
blend.[48]
Solutions and
suspensions[ed
it]
Aqueous
solutions[edit]
See also: Debye–
Hückel
theory and List of
viscosities
§  Aqueous
solutions
Depending on
the solute and
range of
concentration, an
aqueous electrolyte 
solution can have
either a larger or
smaller viscosity
compared with pure
water at the same
temperature and
pressure. For
instance, a 20%
saline (sodium
chloride) solution
has viscosity over
1.5 times that of
pure water,
whereas a
20% potassium
iodide solution has
viscosity about 0.91
times that of pure
water.
An idealized model
of dilute electrolytic
solutions leads to
the following
prediction for the
viscosity  of a
solution:[49]
where  is the
viscosity of the
solvent,  is the
concentration,
and  is a
positive
constant which
depends on
both solvent
and solute
properties.
However, this
expression is
only valid for
very dilute
solutions,
having  less
than 0.1 mol/L.
[50]
 For higher
concentrations,
additional terms
are necessary
which account
for higher-order
molecular
correlations:
where  and  
are fit from
data. In
particular, a
negative
value of  is
able to
account for
the
decrease in
viscosity
observed in
some
solutions.
Estimated
values of
these
constants
are shown
below for
sodium
chloride and
potassium
iodide at
temperature
25 °C (mol
= mole, L
= liter).[49]

 (
S m  (  (
o o m m
l l− o− o−
u 1/2  l1 l2
t L    
e 1/2 L L
) ) )
2

S
o
d
i
u
m

c 0 0 0
h . . .
l 0 0 0
o 0 7 0
r 6 9 8
i 2 3 0
d

(
N
a
C
l)

P 0 − 0
o . 0 .
t 0 . 0
a 0 0 0
s 4 7 0
si 7 5 0
u 5
m
i
o
d
i
d

(
K
I
)

Suspensions
[edit]
In a
suspension
of solid
particles
(e.g. micron
-size
spheres
suspended
in oil), an
effective
viscosity  ca
n be defined
in terms of
stress and
strain
components
which are
averaged
over a
volume
large
compared
with the
distance
between the
suspended
particles,
but small
with respect
to
macroscopi
c
dimensions.
[51]
 Such
suspensions
generally
exhibit non-
Newtonian
behavior.
However,
for dilute
systems in
steady
flows, the
behavior is
Newtonian
and
expressions
for  can be
derived
directly from
the particle
dynamics.
In a very
dilute
system, with
volume
fraction ,
interactions
between the
suspended
particles
can be
ignored. In
such a case
one can
explicitly
calculate
the flow field
around each
particle
independent
ly, and
combine the
results to
obtain . For
spheres,
this results
in the
Einstein
equation:
where  is
the
viscosity
of the
suspendi
ng liquid.
The
linear
depende
nce
on  is a
consequ
ence of
neglectin
g
interparti
cle
interacti
ons. For
dilute
systems
in
general,
one
expects  
to take
the form
wher
e the
coeffi
cient 
 may
depe
nd
on
the
parti
cle
shap
e
(e.g.
sphe
res,
rods,
disks
).[52] E
xperi
ment
al
deter
mina
tion
of
the
preci
se
value
of  is
diffic
ult,
howe
ver:
even
the
predi
ction 
 for
sphe
res
has
not
been
concl
usive
ly
valid
ated,
with
vario
us
expe
rime
nts
findin
g
value
s in
the
rang
e .
This
defici
ency
has
been
attrib
uted
to
diffic
ulty
in
contr
olling
expe
rime
ntal
condi
tions.
[53]

In
dens
er
susp
ensio
ns,  
acqui
res a
nonli
near
depe
nden
ce
on ,
whic
h
indic
ates
the
impo
rtanc
e of
inter
parti
cle
inter
actio
ns.
Vario
us
analy
tical
and
semi
-
empi
rical
sche
mes
exist
for
capt
uring
this
regi
me.
At
the
most
basic
level,
a
term
quad
ratic
in  is
adde
d to :
a
n
d
t
h
e
c
o
e
ffi
ci
e
n

 i
s
fit
fr
o
m

e
x
p
e
ri
m
e
n
t
al
d
a
t
a
o
r
a
p
p
r
o
xi
m
a
t
e
d
fr
o
m

t
h
e
m
ic
r
o
s
c
o
pi
c
t
h
e
o
r
y.
H
o
w
e
v
e
r,
s
o
m
e
a
u
t
h
o
r
s
a
d
vi
s
e
c
a
u
ti
o
n
in
a
p
pl
yi
n
g
s
u
c
h
si
m
pl
e
f
o
r
m
ul
a
s
si
n
c
e
n
o
n
-
N
e
w
t
o
ni
a
n
b
e
h
a
vi
o
r
a
p
p
e
a
r
s
in
d
e
n
s
e
s
u
s
p
e
n
si
o
n
s

f
o
r
s
p
h
e
r
e
s
),
[53]
 
o
r
in
s
u
s
p
e
n
si
o
n
s
o
f
el
o
n
g
a
t
e
d
o
r
fl
e
xi
bl
e
p
a
rt
ic
le
s.
[51]

T
h
e
r
e
is
a
di
st
in
ct
io
n
b
e
t
w
e
e
n
a
s
u
s
p
e
n
si
o
n
o
f
s
ol
id
p
a
rt
ic
le
s,
d
e
s
c
ri
b
e
d
a
b
o
v
e
,
a
n
d
a

e
m
ul
si
o
n
.
T
h
e
la
tt
e
r
is
a
s
u
s
p
e
n
si
o
n
o
f
ti
n
y
d
r
o
pl
e
ts
,
w
hi
c
h
t
h
e
m
s
el
v
e
s
m
a
y
e
x
hi
bi
t
in
t
e
r
n
al
ci
r
c
ul
a
ti
o
n
.
T
h
e
p
r
e
s
e
n
c
e
o
f
in
t
e
r
n
al
ci
r
c
ul
a
ti
o
n
c
a
n
d
e
c
r
e
a
s
e
t
h
e
o
b
s
e
r
v
e
d
e
ff
e
ct
iv
e
vi
s
c
o
si
ty
,
a
n
d
di
ff
e
r
e
n
t
t
h
e
o
r
e
ti
c
al
o
r
s
e
m
i-
e
m
pi
ri
c
al
m
o
d
el
s
m
u
st
b
e
u
s
e
d
.
[54]

A
m
o
r
p
h
o
u
s
m
a
t
e
ri
a
l
s
[
e
di
t]
C
o
m
m
o

gl
a
s

vi
s
c
o
si
ty
c
ur
v
e
s[5
5]

I
n
t
h
e
hi
g
h
a
n
d
lo
w
t
e
m
p
e
r
a
t
u
r
e
li
m
it
s,
vi
s
c
o
u
s
fl
o
w
in 
a
m
o
r
p
h
o
u
s
m
a
t
e
ri
al

(
e
.
g
.
in 
gl
a
s
s
e

a
n
d
m
el
ts
)
[56]

[57]

 
[58]

h
a
s
t
h

A
rr
h
e
ni
u
s
f
o
r
m
:
w
h
er


is
a
re
le
v
a
nt 
a
ct
iv
at
io
n
e
n
er
g
y,
gi
v
e
n
in
te
r
m
s
of
m
ol
e
c
ul
ar
p
ar
a
m
et
er
s; 

is
te
m
p
er
at
ur
e; 

is
th
e
m
ol
ar 
g
a
s
c
o
n
st
a
nt
;
a
n


is
a
p
pr
o
xi
m
at
el
y
a
c
o
n
st
a
nt
.
T
h
e
a
ct
iv
at
io
n
e
n
er
g


ta
k
e
s
a
di
ff
er
e
nt
v
al
u
e
d
e
p
e
n
di
n
g
o
n
w
h
et
h
er
th
e
hi
g
h
or
lo
w
te
m
p
er
at
ur
e
li
m
it
is
b
ei
n
g
c
o
n
si
d
er
e
d:
it
c
h
a
n
g
e
s
fr
o
m
a
hi
g
h
v
al
u

Q

at
lo
w
te
m
p
er
at
ur
e
s
(i
n
th
e
gl
a
s
s
y
st
at
e)
to
a
lo
w
v
al
u

Q

at
hi
g
h
te
m
p
er
at
ur
e
s
(i
n
th
e
li
q
ui
d
st
at
e)
.

Comm
on
logarit
hm of
viscosi
ty
agains
t
tempe
rature
for B2
O3,
showi
ng two
regim
es

F
or
in
te
r
m
e
di
at
e
te
m
p
er
at
ur
e
s, 
 v
ar
ie
s
n
o
nt
ri
vi
al
ly
wi
th
te
m
p
er
at
ur
e
a
n
d
th
e
si
m
pl
e
A
rr
h
e
ni
u
s
fo
r
m
fa
ils
.
O
n
th
e
ot
h
er
h
a
n
d,
th
e
t
w
o-
e
x
p
o
n
e
nt
ia
l
e
q
u
at
io
n
wher
e , , 
,  are
all
const
ants,
provi
des a
good
fit to
exper
iment
al
data
over
the
entire
range
of
temp
eratu
res,
while
at the
same
time
reduc
ing to
the
corre
ct
Arrhe
nius
form
in the
low
and
high
temp
eratu
re
limits.
This
expre
ssion
can
be
motiv
ated
from
vario
us
theor
etical
mode
ls of
amor
phou
s
mater
ials
at the
atomi
c
level.
[57]

A
two-
expo
nenti
al
equat
ion
for
the
visco
sity
can
be
deriv
ed
within
the
Dyre
shovi
ng
mode
l of
super
coole
d
liquid
s,
wher
e the
Arrhe
nius
energ
y
barrie
r is
identi
fied
with
the
high-
frequ
ency 
shear
modu
lus ti
mes
a
chara
cteris
tic
shovi
ng
volu
me.[59]
[60]
 Up
on
speci
fying
the
temp
eratu
re
depe
nden
ce of
the
shear
modu
lus
via
therm
al
expa
nsion
and
via
the
repul
sive
part
of the
inter
mole
cular
poten
tial,
anoth
er
two-
expo
nenti
al
equat
ion is
retrie
ved:[61]
where  d
enotes
the high-
frequenc
y shear
modulus 
of the
material
evaluate
d at a
temperat
ure equal
to
the glass
transition 
temperat
ure ,  is
the so-
called
shoving
volume,
i.e. it is
the
character
istic
volume
of the
group of
atoms
involved
in the
shoving
event by
which an
atom/mol
ecule
escapes
from the
cage of
nearest-
neighbou
rs,
typically
on the
order of
the
volume
occupied
by few
atoms.
Furtherm
ore,  is
the ther
mal
expansio
n coeffici
ent of the
material, 
 is a
paramet
er which
measure
s the
steepnes
s of the
power-
law rise
of the
ascendin
g flank of
the first
peak of
the radial
distributi
on
function,
and is
quantitati
vely
related to
the
repulsive
part of
the intera
tomic
potential.
[61]
 Finally, 
 denotes
the Boltz
mann
constant.
Eddy
viscosi
ty[edit]
In the
study
of turbule
nce in flu
ids, a
common
practical
strategy
is to
ignore
the
small-
scale vor
tices (or 
eddies)
in the
motion
and to
calculate
a large-
scale
motion
with
an effecti
ve viscos
ity, called
the
"eddy
viscosity"
, which
character
izes the
transport
and
dissipatio
n
of energy 
in the
smaller-
scale
flow
(see larg
e eddy
simulatio
n).[62][63] In
contrast
to the
viscosity
of the
fluid
itself,
which
must be
positive
by
the seco
nd law of
thermody
namics,
the eddy
viscosity
can be
negative.
[64][65]

Predi
ction[e
dit]
Because
viscosity
depends
continuo
usly on
temperat
ure and
pressure,
it cannot
be fully
character
ized by a
finite
number
of
experime
ntal
measure
ments.
Predictiv
e
formulas
become
necessar
y if
experime
ntal
values
are not
available
at the
temperat
ures and
pressure
s of
interest.
This
capability
is
important
for
thermop
hysical
simulatio
ns, in
which
the
temperat
ure and
pressure
of a fluid
can vary
continuo
usly with
space
and time.
A similar
situation
is
encounte
red for
mixtures
of pure
fluids,
where
the
viscosity
depends
continuo
usly on
the
concentr
ation
ratios of
the
constitue
nt fluids
For the
simplest
fluids,
such as
dilute
monatom
ic gases
and their
mixtures, 
ab
initio qua
ntum
mechani
cal comp
utations
can
accuratel
y predict
viscosity
in terms
of
fundame
ntal
atomic
constant
s, i.e.,
without
referenc
e to
existing
viscosity
measure
ments.[66] 
For the
special
case of
dilute
helium, u
ncertainti
es in
the ab
initio calc
ulated
viscosity
are two
order of
magnitud
es
smaller
than
uncertain
ties in
experime
ntal
values.[67]
For most
fluids,
such
high-
accuracy
, first-
principle
s
computat
ions are
not
feasible.
Rather,
theoretic
al or
empirical
expressi
ons must
be fit to
existing
viscosity
measure
ments. If
such an
expressi
on is fit
to high-
fidelity
data over
a large
range of
temperat
ures and
pressure
s, then it
is called
a
"referenc
e
correlatio
n" for
that fluid.
Referenc
e
correlatio
ns have
been
publishe
d for
many
pure
fluids; a
few
example
s
are water
, carbon
dioxide, 
ammonia
, benzen
e,
and xeno
n.[68][69][70][71]
[72]
 Many
of these
cover
temperat
ure and
pressure
ranges
that
encompa
ss gas,
liquid,
and supe
rcritical p
hases.
Thermop
hysical
modeling
software
often
relies on
referenc
e
correlatio
ns for
predictin
g
viscosity
at user-
specified
temperat
ure and
pressure.
These
correlatio
ns may
be propri
etary.
Example
s
are REF
PROP[73] (
proprieta
ry)
and Cool
Prop[74] (o
pen-
source).
Viscosity
can also
be
compute
d using
formulas
that
express
it in
terms of
the
statistics
of
individual
particle
trajectori
es.
These
formulas
include
the Gree
n–Kubo
relations 
for the
linear
shear
viscosity
and
the trans
ient time
correlati
on
function 
expressi
ons
derived
by Evans
and
Morriss
in 1988.
[75][31]
 The
advantag
e of
these
expressi
ons is
that they
are
formally
exact
and valid
for
general
systems.
The
disadvan
tage is
that they
require
detailed
knowled
ge of
particle
trajectori
es,
available
only in
computat
ionally
expensiv
e
simulatio
ns such
as molec
ular
dynamic
s. An
accurate
model for
interparti
cle
interactio
ns is also
required,
which
may be
difficult
to obtain
for
complex
molecule
s.[76]

Select
ed
subst
ances[
edit]

In the
University of
Queensland pit
ch 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.[77]

Main
article: L
ist of
viscositie
s
Observe
d values
of
viscosity
vary over
several
orders of
magnitud
e, even
for
common
substanc
es (see
the order
of
magnitud
e table
below).
For
instance,
a 70%
sucrose
(sugar)
solution
has a
viscosity
over 400
times
that of
water,
and
26000
times
that of
air.[78] Mor
e
dramatic
ally, pitch 
has been
estimate
d to have
a
viscosity
230
billion
times
that of
water.[77]
Water[
edit]
The dyna
mic
viscosity 
 of water 
is about
0.89
mPa·s at
room
temperat
ure
(25 °C).
As a
function
of
temperat
ure
in kelvins
, the
viscosity
can be
estimate
d using
the semi-
empirical 
Vogel-
Fulcher-
Tamman
n
equation:
where A =
0.02939
mPa·s, B =
507.88 K,
and C =
149.3 K.
[79]
 Experimen
tally
determined
values of the
viscosity are
also given in
the table
below. The
values at
20 °C are a
useful
reference:
there, the
dynamic
viscosity is
about 1 cP
and the
kinematic
viscosity is
about 1 cSt.

5
7

Air[edit]
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.
Among the
many
possible
approximate
formulas for
the
temperature
dependence
(see Temper
ature
dependence
of viscosity),
one is:[80]
which is
accurate in the
range −20 °C
400 °C. For th
formula to be
valid, the
temperature
must be given
in kelvins;  the
corresponds to
the viscosity in
Pa·s.

Honey being drizz

Other
common
substances
edit]
Order of
magnitude
estimates[e
t]
The following
table illustrate
the range of
viscosity value
observed in
common
substances.
Unless
otherwise note
a temperature
25 °C and a
pressure of 1
atmosphere a
assumed.
The values
listed are
representative
estimates only
as they do not
account for
measurement
uncertainties,
variability in
material
definitions, or
non-Newtonia
behavior.
See
also[edit]
 Dashpot
 Deborah n
 Dilatant
 Herschel–B
fluid
 High visco
 Hypervisco
syndrome
 Intrinsic vis
 Inviscid flo
 Joback
method (es
of liquid vis
from molec
structure)
 Kaye effec
 Microvisco
 Morton num
 Oil pressur
 Quasi-solid
 Rheology
 Stokes flow
 Superfluid
 Viscoplasti
 Viscosity m
mixtures
 Zahn cup

Reference
[edit]
Footnotes[
it]

1. ^ The discuss
follows draws
& Cowling 19
2. ^ a b These ma
highly non-Ne
Citations[ed
1. ^ Symon 197
2. ^ Balescu 19
3. ^ Landau & L
4. ^ Harper, Dou
(n.d.).  "viscou
Etymology
Dictionary. Ar
original on 1 M
Retrieved  19
September 2
5. ^ Mewis & W
p. 19.
6. ^ Streeter, W
1998.
7. ^ Holman 200
8. ^ Incropera e
9. ^ Nič et al. 19
10. ^ a b c Bird, Ste
2007, p. 19.
11. ^ a b Landau &
pp. 44–45.
12. ^ Bird, Stewa
2007, p. 18: T
an alternative
convention, w
reversed here
13. ^ Landau & L
p. 45.
14. ^ a b Balescu 1
15. ^ a b c d Chapm
1970.
16. ^ Millat 1996.
17. ^ a b c d e Bird, S
Lightfoot 2007
18. ^ Schroeder 1
19. ^ Różańska e
pp. 47–55.
20. ^ Trouton 190
21. ^ Mewis & W
pp. 228–230.
22. ^ Kumagai, S
1978, pp. 157
23. ^ Scherer, Pa
Swiatek 1988
24. ^ Hannan 200
25. ^ Kwapiszew
26. ^ McNaught &
1997, poise.
27. ^ Gyllenbok 2
28. ^ ASTM D216
Practice for C
Kinematic Vis
Saybolt Unive
to Saybolt Fu
Viscosity, AS
29. ^ Trachenko
2020.
30. ^ a b Trachenk
2021.
31. ^ a b Evans &
32. ^ a b Bellac, M
Batrouni 2004
33. ^ Chapman &
p. 103.
34. ^ Cercignani
35. ^ Sutherland
531.
36. ^ Bird, Stewa
2007, pp. 25–
37. ^ Chapman &
pp. 235–237.
38. ^ Chapman &
pp. 197, 214–
39. ^ Cramer 201
40. ^ Reid & She
p. 202.
41. ^ a b c Bird, Ste
2007, pp. 29–
42. ^ Reid & She
pp. 203–204.
43. ^ Hildebrand
44. ^ Hildebrand
45. ^ Egelstaff 19
46. ^ Irving & Kirk
pp. 817–829.
47. ^ Reid & She
pp. 206–209.
48. ^ a b Zhmud 20
49. ^ a b Viswanat
50. ^ Abdulagato
Azizov 2006,
51. ^ a b Bird, Stew
2007, pp. 31–
52. ^ Bird, Stewa
2007, p. 32.
53. ^ a b Mueller, L
Mader 2009,
54. ^ Bird, Stewa
2007, p. 33.
55. ^ Fluegel 200
56. ^ Doremus 20
7629.
57. ^ a b Ojovan, T
2007, p. 4151
58. ^ Ojovan & L
pp. 3803–381
59. ^ Dyre, Olsen
1996, p. 2171
60. ^ Hecksher &
61. ^ a b Krausser
Zaccone 2015
62. ^ Bird, Stewa
2007, p. 163.
63. ^ Lesieur 201
64. ^ Sivashinsky
p. 1040.
65. ^ Xie & Levch
p. 045434.
66. ^ Sharipov &
67. ^ Rowland, A
2020.
68. ^ Huber et al.
69. ^ Laesecke &
70. ^ Monogenid
Huber 2018.
71. ^ Avgeri et al
72. ^ Velliadou e
73. ^ "Refprop". N
18 April 2013
the original on
Retrieved  202
74. ^ Bell et al. 2
75. ^ Evans & Mo
76. ^ Maginn et a
77. ^ a b c d Edgew
Parnell 1984,
78. ^ a b c d e Rumb
79. ^ Viswanath &
1989, pp. 714
80. ^ tec-science
25). "Viscosit
gases". tec-s
d from the ori
04-19. Retriev
81. ^ a b c d Fellow
82. ^ Yanniotis, S
Karaburnioti 2
377.
83. ^ a b Koocheki
pp. 596–602.
84. ^ a b Citerne, C
2001, pp. 86–
85. ^ Kestin, Kha
1977.
86. ^ Assael et a
87. ^ Kestin, Ro &
1972.
88. ^ Rosenson,
Uretz 1996.
89. ^ Zhao et al.
90. ^ Sagdeev et
91. ^ Walzer, He
Baumgardner

Sources[ed
 Abdulag
ov,
Ilmutdin
M.;
Zeinalov
Adelya B
Azizov,
Nazim D
(2006).
"Experim
ntal
viscosity
B-
coefficie
s of
aqueous
LiCl
solutions
Journal o
Molecula
Liquids. 
6 (1–3):
75–88.  d
i:10.1016
molliq.20
5.10.006
SSN  016
-7322.
 Assael, M
J.; et al.
(2018). "
eference
Values
and
Referenc
Correlati
s for the
Thermal
Conduct
ty and
Viscosity
of
Fluids". 
urnal of
Physical
and
Chemica
Referenc
Data. 47
2):
021501. 
bcode:20
8JPCRD
47b1501
doi:10.10
3/1.5036
25. ISSN
0047-
2689. PM
C  64633
0.  PMID
0996494
 Avgeri, S
Assael, M
J.; Hube
M. L.;
Perkins,
A. (2014
"Referen
e
Correlati
of the
Viscosity
of
Benzene
from the
Triple
Point to
675 K an
up to 300
MPa".  Jo
rnal of
Physical
and
Chemica
Referenc
Data. AIP
Publishin
43  (3):
033103. 
bcode:20
4JPCRD
43c3103
doi:10.10
3/1.4892
35. ISSN
0047-
2689.
 Balescu,
Radu (1
5). Equil
ium and
Non-
Equilibriu
m
Statistica
Mechani
. John
Wiley &
Sons. IS
N  978-0-
471-
04600-4
Archived
om the
original o
2020-03
16.
Retrieve
2019-09
18.
 Bell, Ian
H.;
Wronski,
Jorrit;
Quoilin,
Sylvain;
Lemort,
Vincent
(2014-01
27). "Pu
and
Pseudo-
pure Flu
Thermop
ysical
Property
Evaluatio
and the
Open-
Source
Thermop
ysical
Property
Library
CoolProp
Industria
&
Enginee
g
Chemist
Researc
America
Chemica
Society
(ACS). 5
(6): 2498
2508. do
10.1021/
4033999
SSN  088
-5885.  P
C  39446
5.  PMID
4623957
 Bellac,
Michael;
Mortessa
ne,
Fabrice;
Batrouni
G. Georg
(2004). E
uilibrium
and Non
Equilibriu
m
Statistica
Thermod
namics.
Cambrid
Universit
Press.  IS
N  978-0-
521-
82143-8

 Bird, R.
Byron;
Stewart,
Warren E
Lightfoot
Edwin N
(2007). T
ansport
Phenom
a (2nd  e
). John
Wiley &
Sons,
Inc. ISBN
978-0-47
11539-8
Archived
om the
original o
2020-03
02.
Retrieve
2019-09
18.

 Bird, R.
Bryon;
Armstron
Robert C
Hassage
Ole
(1987), D
namics o
Polymer
Liquids,
Volume
Fluid
Mechani
(2nd ed.
John Wil
& Sons

 Cercigna
,
Carlo (1
5). Theo
and
Applicati
of the
Boltzman
Equation
Elsevier.
SBN  978
0-444-
19450-3

 Chapma
Sydney; 
owling,
T.G. (19
). The
Mathema
cal Theo
of Non-
Uniform
Gases  (3
d ed.).
Cambrid
Universit
Press.  IS
N  97805
1075770
 Citerne,
Guillaum
P.;
Carreau,
Pierre J.
Moan,
Michel
(2001).
"Rheolog
al
propertie
of peanu
butter". R
eologica
Acta.  40
1): 86–
96. doi:1
1007/s00
9700001
0.  S2CID
9455582

 Cramer,
M.S.
(2012). "
umerical
estimate
for the
bulk
viscosity
ideal
gases". 
ysics of
Fluids. 2
(6):
066102–
066102–
23. Bibc
e:2012P
l...24f610
C. doi:10
063/1.47
9611. hd
10919/47
46. Arch
ed  from
the origin
on 2022-
02-15.
Retrieve
2020-09
19.

 Doremus
R.H.
(2002).
"Viscosit
of
silica".  J
Appl.
Phys.  92
12): 761
7629. Bi
ode:2002
AP....92.
619D.  do
10.1063/
1515132
 Dyre, J.C
Olsen, N
B.;
Christen
n, T.
(1996). "
ocal
elastic
expansio
model fo
viscous-
flow
activatio
energies
of glass-
forming
molecula
liquids". 
hysical
Review
B. 53 (5
2171–
2174. Bi
ode:1996
hRvB..53
2171D.  d
:10.1103
hysRevB
3.2171. 
MID 998
702.
 Edgewor
, R.;
Dalton,
B.J.;
Parnell,
(1984). "
he pitch
drop
experime
t".  Europ
an Journ
of
Physics.
(4): 198–
200. Bib
de:1984E
Ph....5..1
8E.  doi:1
1088/014
-
0807/5/4
03. S2C
2507695
9.  Archiv
d from th
original o
2013-03
28.
Retrieve
2009-03
31.
 Egelstaff
P. A.
(1992). A
Introduct
n to the
Liquid
State (2n
ed.).
Oxford
Universit
Press.  IS
N  978-0-
19-
851012-
 Evans,
Denis J.;
Morriss,
Gary P.
(2007). S
atistical
Mechani
of
Nonequi
rium
Liquids.
ANU
Press.  IS
N  97819
1313226
JSTOR 
t24h99q.
rchived  f
m the
original o
2022-01
10.
Retrieve
2022-01
10.
 Evans,
Denis J.;
Morriss,
Gary P.
(October
15, 1988
"Transie
time-
correlatio
functions
and the
rheology
of
fluids". P
ysical
Review
A. 38 (8
4142–
4148. Bi
ode:1988
hRvA..38
4142E. d
:10.1103
hysRevA
8.4142. 
MID 990
865.
 Fellows,
J.
(2009). F
od
Processi
g
Technolo
y:
Principle
and
Practice 
rd  ed.).
Woodhe
.  ISBN  9
8-
1845692
62.

 Fluegel,
Alexande
(2007). "
scosity
calculatio
of
glasses"
Glasspro
erties.co
Archived
om the
original o
2010-11
27.
Retrieve
2010-09
14.
 Gibbs,
Philip
(January
1997).  "I
glass
liquid or
solid?".  m
ath.ucr.e
u.  Archiv
d from th
original o
29 March
2007.
Retrieve
19
Septemb
2019.
 Gyllenbo
Jan (201
.
"Encyclo
aedia of
Historica
Metrolog
Weights,
and
Measure
Volume
1". Ency
opaedia
Historica
Metrolog
Weights,
and
Measure
Vol.  1.
Birkhäus
.  ISBN  9
8331957
988.
 Hannan,
Henry
(2007). T
chnician
Formula
n
Handboo
for
Industria
and
Househo
Cleaning
Products
Waukesh
,
Wiscons
Kyral LL
p.  7.  ISB
978-0-
6151-
5601-9.
 Heckshe
Tina; Dy
Jeppe C
(2015-01
01). "A
review o
experime
ts testing
the
shoving
model". 
urnal of
Non-
Crystallin
Solids. 7
IDMRCS
Relaxatio
in
Complex
Systems
407: 14–
22. Bibc
e:2015JN
CS..407.
14H. doi
0.1016/j.
oncrysol
014.08.0
6.  ISSN 
022-309
Archived
om the
original o
2022-02
15.
Retrieve
2021-10
17.
 Hildebra
, Joel
Henry  (1
77). Visc
sity and
Diffusivit
A
Predictiv
Treatme
John Wil
&
Sons. IS
N  978-0-
471-
03072-0

 Holman,
Jack Phi
(2002). H
at
Transfer
McGraw
Hill.  ISBN
978-0-07
112230-
Archived
om the
original o
2020-03
15.
Retrieve
2019-09
18.
 Huber, M
L.;
Perkins,
A.;
Laeseck
A.; Frien
D. G.;
Sengers
J. V.;
Assael, M
J.; Metax
I. N.;
Vogel, E
Mareš, R
Miyagaw
K. (2009
"New
Internatio
al
Formula
n for the
Viscosity
of
H2O".  Jo
rnal of
Physical
and
Chemica
Referenc
Data. AIP
Publishin
38  (2):
101–125
Bibcode:
009JPCR
D..38..10
H. doi:10
063/1.30
8050. IS
N  0047-
2689.
 Incroper
Frank P.
et  al.
(2007). F
ndament
s of Hea
and Mas
Transfer
Wiley.  IS
N  978-0-
471-
45728-2
Archived
om the
original o
2020-03
11.
Retrieve
2019-09
18.
 Irving,
J.H.;  Kirk
ood, Joh
G.  (1949
"The
Statistica
Mechani
l Theory
Transpo
Processe
. IV. The
Equation
of
Hydrody
mics".  J.
Chem.
Phys.  18
6): 817–
829. doi
0.1063/1
747782.

 Kestin, J
Ro, S. T
Wakeha
W. A.
(1972).
"Viscosit
of the
Noble
Gases in
the
Tempera
re Range
25–
700°C". 
he Journ
of
Chemica
Physics.
6 (8):
4119–
4124. Bi
ode:1972
ChPh..56
4119K. d
:10.1063
.1677824
ISSN  00
-9606.
 Kestin, J
Khalifa,
H.E.;
Wakeha
W.A.
(1977).
"The
viscosity
five
gaseous
hydrocar
ons".  Th
Journal o
Chemica
Physics.
6 (3):
1132. Bi
ode:1977
ChPh..66
1132K. d
:10.1063
.434048.
 Koochek
Arash;
et  al.
(2009).
"The
rheologic
propertie
of ketchu
as a
function
different
hydrocol
ds and
tempera
e". Intern
tional
Journal o
Food
Science
Technolo
y. 44 (3)
596–602
doi:10.11
1/j.1365-
2621.200
.01868.x
 Krausse
J.;
Samwer
K.;
Zaccone
A.
(2015). "
teratomic
repulsion
softness
directly
controls
the fragil
of
supercoo
d metalli
melts".  P
oceeding
of the
National
Academ
of
Sciences
of the
USA.  11
(45):
13762–
13767.  a
Xiv:1510
8117. Bi
ode:2015
NAS..11
3762K. d
:10.1073
pnas.150
741112. 
MC 4653
54. PMID
2650420
 Kumaga
Naoichi;
Sasajima
Sadao; I
Hidebum
(15
February
1978).  "L
ng-term
Creep of
Rocks:
Results
with Larg
Specime
s Obtain
in about
20 Years
and Tho
with Sma
Specime
s in abou
3
Years".  J
urnal of
the
Society o
Materials
Science
(Japan).
7 (293):
157–161
NAID 11
0022993
7.  Archiv
d from th
original o
2011-05
21.
Retrieve
2008-06
16.
 Kwapisz
wska,
Karina;
Szczepa
ki,
Krzyszto
Kalwarcz
k, Tomas
Michalsk
Bernade
Patalas-
Krawczy
Paulina;
Szymańs
, Jędrzej
Andrysze
ski,
Tomasz;
Iwan,
Michalin
Duszyńs
Jerzy;
Hołyst,
Robert
(2020). "
anoscale
Viscosity
of
Cytoplas
Is
Conserv
in Huma
Cell
Lines".  T
e Journa
of Physic
Chemist
Letters. 
(16):
6914–
6920. do
10.1021/
cs.jpclett
c01748. 
MC 7450
58. PMID
3278720
 Laeseck
Arno;
Muzny,
Chris D.
(2017). "
eference
Correlati
for the
Viscosity
of Carbo
Dioxide"
ournal of
Physical
and
Chemica
Referenc
Data. AIP
Publishin
46  (1):
013107. 
bcode:20
7JPCRD
46a3107
doi:10.10
3/1.4977
29. ISSN
0047-
2689. PM
C  55146
2.  PMID
8736460
 Landau,
D.;  Lifsh
,
E.M.  (19
7). Fluid
Mechani
(2nd ed.
Elsevier.
SBN  978
0-08-
057073-
Archived
om the
original o
2020-03
21.
Retrieve
2019-09
18.
 Maginn,
Edward
Messerly
Richard
A.;
Carlson,
Daniel J.
Roe,
Daniel R
Elliott, J.
Richard
(2019). "
est
Practices
for
Computi
Transpo
Propertie
1. Self-
Diffusivit
and
Viscosity
from
Equilibriu
m
Molecula
Dynamic
[Article
v1.0]".  L
ng Journ
of
Computa
onal
Molecula
Science.
Universit
of
Colorado
at
Boulder.
(1).  doi:1
33011/liv
coms.1.1
324. ISS
2575-
6524. S2
ID 10435
320.
 Monogen
dou, S. A
Assael, M
J.; Hube
M. L.
(2018). "
eference
Correlati
for the
Viscosity
of
Ammoni
from the
Triple
Point to
725 K an
up to 50
MPa".  Jo
rnal of
Physical
and
Chemica
Referenc
Data. AIP
Publishin
47  (2):
023102. 
bcode:20
8JPCRD
47b3102
.  doi:10.
63/1.503
724. ISS
0047-
2689. PM
C  65128
9.  PMID
1092958
 Lesieur,
Marcel
(2012). T
rbulence
Fluids:
Stochast
and
Numeric
Modellin
Springer
SBN  978
94-009-
0533-7. 
chived  fr
m the
original o
2020-03
14.
Retrieve
2018-11
30.
 Mewis,
Jan;
Wagner,
Norman
(2012). C
lloidal
Suspens
n
Rheolog
Cambrid
Universit
Press.  IS
N  978-0-
521-
51599-3
Archived
om the
original o
2020-03
14.
Retrieve
2018-12
10.
 McNaug
A. D.;
Wilkinso
A. (1997
"poise". 
PAC.
Compen
um of
Chemica
Termino
gy (the
"Gold
Book"). S
J. Chalk
(2nd ed.
Oxford:
Blackwe
Scientific
doi:10.13
1/goldbo
.  ISBN  0
9678550
9-8.
 Millat,
Jorgen
(1996). T
ansport
Propertie
of Fluids
Their
Correlati
,
Predictio
and
Estimatio
.
Cambrid
:
Cambrid
Universit
Press.  IS
N  978-0-
521-
02290-3
OCLC 6
204060.
 Mueller,
S.;
Llewellin
E. W.;
Mader, H
M.
(2009). "
he
rheology
of
suspens
ns of sol
particles
Proceed
gs of the
Royal
Society A
Mathema
cal,
Physical
and
Enginee
g
Sciences
466  (211
: 1201–
1228. do
10.1098/
pa.2009.
445. ISS
1364-
5021.
 Nič,
Miloslav;
et  al., ed
(1997).
"dynamic
viscosity
η". IUPA
Compen
um of
Chemica
Termino
gy. Oxfo
Blackwe
Scientific
Publicati
s. doi:10
351/gold
ook.  ISB
978-0-
9678550
9-7.
 Ojovan,
M.I.; Lee
W.E.
(2004).
"Viscosit
of netwo
liquids
within
Doremus
approach
J. Appl.
Phys.  95
7): 3803
3810. Bi
ode:2004
AP....95.
803O. d
10.1063/
1647260
 Ojovan,
M.I.;
Travis, K
P.; Hand
R.J.
(2007). "
hermody
amic
paramet
s of bond
in glassy
materials
from
viscosity
tempera
e
relations
ps"  (PDF
J. Phys.:
Condens
Matter. 1
(41):
415107. 
bcode:20
7JPCM..
9O5107O
doi:10.10
8/0953-
8984/19/
1/415107
PMID  28
92319.  S
CID  247
512. Arc
ved  (PDF
from the
original o
2018-07
25.
Retrieve
2019-09
27.
 Plumb,
Robert C
(1989). "
ntique
windowp
nes and
the flow
supercoo
d
liquids". 
urnal of
Chemica
Educatio
66  (12):
994. Bib
de:1989J
hEd..66.
94P.  doi
0.1021/e
066p994
Archived
from the
original  o
2005-08
26.
Retrieve
2013-12
25.
 Rapapor
D.C.
(2004). T
e Art of
Molecula
Dynamic
Simulatio
(2nd ed.
Cambrid
Universit
Press.  IS
N  97805
1825689
Archived
om the
original o
2018-06
25.
Retrieve
2022-01
10.
 Reid,
Robert C
Sherwoo
Thomas
(1958). T
e
Propertie
of Gases
and
Liquids.
McGraw
Hill.
 Reif, F.
(1965), F
ndament
s of
Statistica
and
Thermal
Physics,
McGraw
Hill. An
advance
treatmen
 Rosenso
R S;
McCorm
k, A;
Uretz, E
(1996-08
01). "Dis
bution of
blood
viscosity
values a
biochem
al
correlate
in health
adults". 
nical
Chemist
Oxford
Universit
Press
(OUP).  4
(8): 1189
1195. do
10.1093/
nchem/4
8.1189. 
SN 0009
9147. PM
D  86975
5.
 Rowland
Darren; A
Ghafri,
Saif Z. S
May, Eri
F. (2020
03-01). "
ide-
Ranging
Referenc
Correlati
s for Dilu
Gas
Transpo
Propertie
Based o
Ab Initio
Calculati
ns and
Viscosity
Ratio
Measure
ents". Jo
nal of
Physical
and
Chemica
Referenc
Data. AIP
Publishin
49  (1):
013101. 
bcode:20
0JPCRD
49a3101
doi:10.10
3/1.5125
00. ISSN
0047-
2689. S2
ID 21379
612.
 Różańsk
S.;
Różańsk
J.;
Ochowia
M.;
Mitkowsk
P. T.
(2014). "
xtension
viscosity
measure
ents of
concentr
ed
emulsion
with the
use of th
opposed
nozzles
device"  (
DF).  Bra
ian Journ
of
Chemica
Enginee
g.  31  (1)
47–55.  d
i:10.1590
S0104-
6632201
0001000
6.  ISSN 
104-663
Archived
PDF) fro
the origin
on 2020-
05-08.
Retrieve
2019-09
19.
 Rumble,
John R.,
ed.
(2018). C
RC
Handboo
of
Chemist
and
Physics 
9th ed.).
Boca
Raton, F
CRC
Press.  IS
N  978-
1138561
32.
 Sagdeev
Damir;
Gabitov,
Il'giz;
Isyanov,
Chingiz;
Khairutd
ov, Vene
Farakho
Mansur;
Zaripov,
Zufar;
Abdulag
ov,
Ilmutdin
(2019-04
22).
"Densitie
and
Viscositi
of Oleic
Acid at
Atmosph
ic
Pressure
Journal o
the
America
Oil
Chemist
Society.
Wiley.  96
6): 647–
662. doi
0.1002/a
cs.12217
ISSN  00
-021X. S
CID  150
6106.
 Scherer,
George
W.;
Pardene
Sandra A
Swiatek,
Rose M.
(1988).
"Viscoela
icity in
silica
gel". Jou
al of Non
Crystallin
Solids. 1
7 (1):
14. Bibc
e:1988JN
CS..107.
14S.  doi
0.1016/0
22-
3093(88
0086-5.
 Schroed
Daniel V
(1999). A
Introduct
n to
Thermal
Physics.
Addison
Wesley. 
BN 978-
201-
38027-9
Archived
om the
original o
2020-03
10.
Retrieve
2018-11
30.
 Sharipov
Felix;
Benites,
Victor J.
(2020-07
01).
"Transpo
coefficie
s of mult
compone
t mixture
of noble
gases
based on
ab initio
potential
Viscosity
and
thermal
conducti
y".  Phys
s of Fluid
AIP
Publishin
32  (7):
077104. 
Xiv:2006
8687. Bi
ode:2020
hFl...32g
104S.  do
10.1063/
0016261
SSN  107
-6631.  S
CID  219
8359.
 Sivashin
y, V.;
Yakhot,
(1985).
"Negativ
viscosity
effect in
large-sca
flows".  T
e Physic
of
Fluids. 2
(4):
1040. Bi
ode:1985
hFl...28.1
40S.  doi
0.1063/1
65025.
 Streeter,
Victor
Lyle;
Wylie, E
Benjamin
Bedford,
Keith W.
(1998). F
id
Mechani
.
WCB/Mc
raw
Hill.  ISBN
978-0-07
062537-
Archived
om the
original o
2020-03
16.
Retrieve
2019-09
18.
 Sutherla
,
William (
893).  "LI
The
viscosity
gases an
molecula
force"  (P
F).  The
London,
Edinburg
and Dub
Philosop
cal
Magazin
and
Journal o
Science.
6 (223):
507–531
doi:10.10
0/147864
9308620
08. ISSN
1941-
5982. Ar
ived  (PD
from the
original o
2019-07
20.
Retrieve
2019-09
18.
 Symon,
Keith R.
(1971). M
chanics 
rd  ed.).
Addison-
Wesley. 
BN 978-
201-
07392-8
Archived
om the
original o
2020-03
11.
Retrieve
2019-09
18.
 Trachen
, K.;
Brazhkin
V. V.
(2020-04
22). "Min
mal
quantum
viscosity
from
fundame
al physic
constant
Science
Advance
America
Associat
n for the
Advance
ent of
Science
(AAAS). 
(17):
eaba374
arXiv:19
.06711. 
bcode:20
0SciA....
3747T. d
:10.1126
sciadv.a
3747. IS
N  2375-
2548. PM
C  71824
0.  PMID
2426470
 Trachen
, Kostya;
Brazhkin
Vadim V
(2021-12
01). "The
quantum
mechani
of
viscosity
PDF).  Ph
sics
Today.
AIP
Publishin
74  (12):
66–67.  B
code:202
PhT....74
66T. doi
0.1063/p
3.4908. 
SN 0031
9228. S2
ID 24483
744. Arc
ved  (PDF
from the
original o
2022-01
10.
Retrieve
2022-01
10.
 Trouton,
Fred. T.
(1906). "
n the
Coefficie
of Viscou
Traction
and Its
Relation
that of
Viscosity
Proceed
gs of the
Royal
Society A
Mathema
cal,
Physical
and
Enginee
g
Sciences
77  (519)
426–440
Bibcode:
906RSP
A..77..42
T.  doi:10
098/rspa
906.0038
ISSN  13
-5021.
 Velliadou
Danai;
Tasidou,
Katerina
A.;
Antoniad
,
Konstan
os D.;
Assael,
Marc J.;
Perkins,
Richard
A.; Hube
Marcia L
(2021-03
25). "Re
ence
Correlati
for the
Viscosity
of Xenon
from the
Triple
Point to
750 K an
up to 86
MPa".  In
rnational
Journal o
Thermop
ysics.
Springer
Science
and
Business
Media
LLC.  42 
):
74. Bibc
e:2021IJ
...42...74
doi:10.10
7/s10765
021-
02818-9
SSN  019
-928X. P
C  83561
9.  PMID
4393314
 Viswana
D.S.;
Nataraja
G.
(1989). D
ta Book
the
Viscosity
of Liquid
Hemisph
e
Publishin
Corporat
n.  ISBN 
89116-
778-1.
 Viswana
Dabir S.;
et  al.
(2007). V
cosity of
Liquids:
Theory,
Estimatio
,
Experim
t, and
Data.
Springer
SBN  978
1-4020-
5481-5.
 Walzer,
Uwe;
Hendel,
Roland;
Baumga
ner, John
(n.d.),  M
tle
Viscosity
and the
Thicknes
of the
Convect
e
Downwe
ngs,
archived
from the
original  o
2007-06
11
 Xie, Hon
Yi;
Levchen
, Alex (2
January
2019).
"Negativ
viscosity
and eddy
flow of th
imbalanc
d electro
hole liqu
in
graphen
Phys. Re
B. 99 (4
045434. 
Xiv:1807
4770. Bi
ode:2019
hRvB..99
5434X. d
:10.1103
PhysRev
.99.0454
4.  S2CID
5179270
 Yanniotis
S.; Skalt
S.;
Karaburn
ti, S.
(Februar
2006).
"Effect o
moisture
content o
the
viscosity
honey at
different
tempera
es".  Jou
al of Foo
Enginee
g.  72  (4)
372–377
doi:10.10
6/j.jfoode
g.2004.1
017.
 Zhao,
Mengjing
Wang,
Yong;
Yang,
Shufeng
Li,
Jingshe;
Liu, Wei;
Song,
Zhaoqi
(2021). "
ow
behavior
and heat
transfer
molten
steel in a
two-stran
tundish
heated b
plasma".
ournal of
Materials
Researc
and
Technolo
y. Elsevi
BV. 13:
561–572
doi:10.10
6/j.jmrt.2
21.04.06
ISSN  22
-7854.  S
CID  236
7034.
 Zhmud,
Boris
(2014). "
scosity
Blending
Equation
(PDF).
Lube-
Tech:93.
ube.
No. 121.
pp. 22–2
Archived
PDF) fro
the origin
on 2018-
12-01.
Retrieve
2018-11
30.
 "NIST
Referenc
Fluid
Thermod
namic an
Transpo
Propertie
Databas
(REFPR
P):
Version
10". Nist
2018-01
01. Arch
ed  from
the origin
on 2021-
12-16.
Retrieve
2021-12
23.
 tec-
science
(2020-03
25). "Vis
sity of
liquids a
gases". 
c-scienc
Archived
om the
original o
2020-04
19.
Retrieve
2020-05
07.

External
links[edit]

Look up viscosity
Wiktionary, the free
dictionary.

Wikisource has the
of The New Studen
Reference Work ar
"Viscosity of Liqu

Wikiquote has quo


related to Viscosit

 Viscosity -
Feynman L
on Physics
 Fluid prope
high accur
calculation
viscosity fo
frequently
encountere
liquids and
 Fluid
Characteri
Chart – a t
viscosities
vapor pres
various flui
 Gas Dynam
Toolbox –
coefficient
viscosity fo
mixtures o
 Glass Visc
Measurem
viscosity
measurem
viscosity u
fixpoints, g
viscosity ca
 Kinematic
Viscosity –
conversion
kinematic a
dynamic vi
 Physical
Characteri
Water – a
water visco
function of
temperatur
 Calculation
temperatur
dependent
viscosities
common
componen
 Artificial vis
 Viscosity o
Dynamic a
Kinematic,
Engineers
 v

 t

 e
Branches of physics

 Pure

Divisions  Applied 
o Engineering

 Experimental

Approaches  Theoretical 
o Computational

Classical  Classical mechanics 


o Newtonian

o Analytical

o Celestial

o Continuum

 Acoustics

 Classical electromagnetism
 Classical optics 
o Ray

o Wave

 Thermodynamics 
o Statistical

o Non-equilibrium

 Relativistic mechanics 
o Special

o General

 Nuclear physics

 Quantum mechanics

Modern  Particle physics

 Atomic, molecular, and optical physics 


o Atomic

o Molecular

o Modern optics

 Condensed matter physics

 Astrophysics

 Atmospheric physics

 Biophysics

 Chemical physics

 Geophysics
Interdisciplinary
 Materials science

 Mathematical physics

 Medical physics

 Ocean physics

 Quantum information science

 History of physics

 Nobel Prize in Physics

Related  Philosophy of physics

 Physics education

 Timeline of physics discoveries

Authority control: National   Spain

 France

 BnF data

 Germany
 Israel

 United States

 Japan

 Czech Republic
Portal:

  Physics
Categories: 
 Viscosity
 Aerodynam
 Fluid dyna
 This page was
edited on 20 A
2023, at 15:35
 Text is availab
under the Crea
Commons Attr
ShareAlike Lic
3.0; additional
may apply. By
this site, you a
the Terms of
Use and Priva
Policy. Wikipe
a registered
trademark of
the Wikimedia
Foundation, In
non-profit
organization.
 Privacy policy

 About Wikiped

 Disclaimers

 Contact Wikip

 Mobile view

 Developers

 Statistics

 Cookie statem

Toggle limited
content width
C L-36-A-90 (ASTM D 4741) © 2019 Chevron Marine Products LLC. All rights reserved. All trademarks
are the property of Chevron Intellectual Property LLC or their respective owners.

You might also like