Viscosity


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.
Viscosity can be conceptualized as quantifying the internal 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 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; such fluids are technically said to be viscous or viscid. A fluid with a high viscosity, such as pitch, may appear to be a solid.

Etymology

The word "viscosity" is derived from the Latin viscum. Viscum also referred to a viscous glue derived from mistletoe berries.

Definition

Simple definition

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. 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 . If the speed of the top plate is low enough, then in steady state the fluid particles move parallel to it, and their speed varies from at the bottom to at the top. 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 . 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. 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, 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. However, the Greek letter eta is also used by chemists, physicists, and the IUPAC. 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.

General definition

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. 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. 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 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. This equation can be thought of as a generalized form of Newton's law of viscosity.
The bulk 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. 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, defined as the ratio of the viscosity to the density of the fluid. It is usually denoted by the Greek letter nu and has dimension :
Consistent with this nomenclature, the viscosity is frequently called the dynamic viscosity or absolute viscosity, and has units force × time/area.

Momentum transport

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 diffusivity characterizes mass transport. 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. The fact that mass, momentum, and energy 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

Newton's law of viscosity is not a fundamental law of nature, but rather a constitutive equation 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:
Trouton's ratio is the ratio of extensional viscosity to shear viscosity. For a Newtonian fluid, the Trouton ratio is 3. Shear-thinning liquids are very commonly, but misleadingly, described as thixotropic.
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.
However, many liquids will briefly react like elastic solids when subjected to sudden stress. Conversely, many "solids" will flow like liquids, albeit very slowly, even under arbitrarily small stress. Such materials are therefore best described as possessing both elasticity and viscosity ; that is, being viscoelastic.
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.

Measurement

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. The fluids without a constant viscosity 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 through the conversion equations.
Also used in coatings, a Stormer viscometer uses load-based rotation in order to determine viscosity. The viscosity is reported in Krebs units, 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.
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 newton-second per square meter, also frequently expressed in the equivalent forms pascal-second and kilogram per meter per second. The CGS unit is the poise, named after Jean Léonard Marie Poiseuille. It is commonly expressed, particularly in ASTM standards, as centipoise, because centipoise is equal to the SI millipascal seconds.
The SI unit of kinematic viscosity is square meter per second, whereas the CGS unit for kinematic viscosity is the stokes, named after Sir George Gabriel Stokes. In U.S. usage, stoke is sometimes used as the singular form. The submultiple centistokes is often used instead, 1 cSt = 1 mm2·s−1 = 10−6 m2·s−1.
The most frequently used systems of US customary, or Imperial, units are the British Gravitational and English Engineering. In the BG system, dynamic viscosity has units of pound-seconds per square foot, and in the EE system it has units of pound-force-seconds per square foot. Note that 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 is defined by Newton's Second Law, whereas in the EE system the units of force and mass are defined independently through the Second Law using the gc proportionality constant gc.
Kinematic viscosity has units of square feet per second in both the BG and EE systems.
Nonstandard units include the reyn, a British unit of dynamic viscosity. 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, sometimes called the rhe. Fluidity is seldom used in engineering practice.
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. Other abbreviations such as SSU or SUV 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

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 1988. 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.
Momentum transport in gases is generally mediated by discrete molecular collisions, and in liquids by attractive forces which bind molecules close together. Because of this, the dynamic viscosities of liquids are typically much larger than those of gases.

Pure gases

Consider a dilute gas moving parallel to the -axis with velocity that depends only on the coordinate. To simplify the discussion, the gas is assumed to have uniform temperature and density.
Under these assumptions, the velocity of a molecule passing through is equal to whatever velocity that molecule had when its mean free path began. Because is typically small compared with macroscopic scales, the average velocity of such a molecule has the form
where is a numerical constant on the order of. Now, because half the molecules on either side are moving towards, and doing so on average with half the average moleculer speed, the momentum flux from either side is
The net momentum flux at is the difference of the two:
According to the definition of viscosity, this momentum flux should be equal to, which leads to
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 interesting 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. Note that this behavior runs counter to common intuition regarding liquids, for which viscosity typically decreases with temperature.
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.

Chapman–Enskog theory

A technique developed by Sydney Chapman and David Enskog in the early 1900s allows a more refined calculation of. It is based on the Boltzmann equation, which provides a systematic statistical description of a dilute gas in terms of intermolecular interactions. As such, their technique allows accurate calculation of for more realistic molecular models, such as those incorporating intermolecular attraction rather than just hard-core repulsion.
It turns out that a more realistic modeling of interactions is essential for accurate prediction of the temperature dependence of, which experiments show increases more rapidly than the trend predicted for rigid elastic spheres. 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, which describes rigid elastic spheres with weak mutual attraction. In such a case, the attractive force can be treated perturbatively, which leads to a particularly 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. If is known from experiments at and at least one other temperature, then can be calculated. It turns out that expressions for obtained in this way are accurate for a number of gases over a sizable range of temperatures. On the other hand, argue that this success does not imply that molecules actually interact according to the Sutherland model. Rather, they interpret the prediction for as a simple interpolation which is valid for some gases over fixed ranges of temperature, but otherwise does not provide a picture of intermolecular interactions which is fundamentally correct and general. Slightly more sophisticated models, such as the Lennard-Jones potential, may provide a better picture, 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.

Bulk viscosity

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 in negligible, but is nonzero for a gas like carbon dioxide, whose molecules possess both rotational and vibrational energy.

Pure liquids

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 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.
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. 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
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 widespread and accurate empirical relation
where and are constants fit from data. On the other hand, several authors express caution with respect to this model.
Errors as large as 30% can be encountered using equation, compared with fitting equation to experimental data. More fundamentally, the physical assumptions underlying equation have been criticized. It has also been argued that the exponential dependence in equation does not necessarily describe experimental observations more accurately than simpler, non-exponential expressions.
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. 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 appear to be the only consistently reliable means of calculating viscosity in liquids.

Mixtures and blends

Gaseous mixtures

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. 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

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 useful 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.
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, besides those stemming directly from molecular theory.

Solutions and suspensions

Aqeuous 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 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:
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. 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.
Solute
Sodium chloride 0.00620.07930.0080
Potassium iodide 0.0047−0.07550.0000

Suspensions

In a suspension of solid particles, an effective viscosity can 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 macroscopic dimensions. 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 independently, and combine the results to obtain. For spheres, this results in the Einstein equation:
where is the viscosity of the suspending liquid. The linear dependence on is a direct consequence of neglecting interparticle interactions; in general, one will have
where the coefficient may depend on the particle shape. Experimental determination of the precise value of is difficult, however: even the prediction for spheres has not been conclusively validated, with various experiments finding values in the range. This deficiency has been attributed to difficulty in controlling experimental conditions.
In denser suspensions, acquires a nonlinear dependence on, which indicates the importance of interparticle interactions. Various analytical and semi-empirical schemes exist for capturing this regime. At the most basic level, a term quadratic in is added to :
and the coefficient is fit from experimental data or approximated from the microscopic theory. In general, however, one should be cautious in applying such simple formulas since non-Newtonian behavior appears in dense suspensions, or in suspensions of elongated or flexible particles.
There is a distinction between a suspension of solid particles, described above, and an emulsion. The latter is a suspension of tiny droplets, which themselves may exhibit internal circulation. The presence of internal circulation can noticeably decrease the observed effective viscosity, and different theoretical or semi-empirical models must be used.

Amorphous materials

In the high and low temperature limits, viscous flow in amorphous materials has the Arrhenius form:
where is a relevant activation energy, given in terms of molecular parameters; is temperature; is the molar gas constant; and is approximately a constant. The activation energy takes a different value depending on whether the high or low temperature limit is being considered: it changes from a high value at low temperatures to a low value at high temperatures.
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.
A two-exponential equation for the viscosity can be derived within the Dyre shoving model of supercooled liquids, where the Arrhenius energy barrier is identified with the high-frequency shear modulus times a characteristic shoving volume. Upon specifying the temperature dependence of the shear modulus via thermal expansion and via the repulsive part of the intermolecular potential, a two-exponential equation is retrieved as in the Krausser-Samwer-Zaccone equation. The parameters in the KSZ equation are expressed in terms of physical quantities such as the repulsive part of the interatomic or intermolecular interaction, the thermal expansion coefficient and the glass transition temperature of the system.
The KSZ equation reads as:
where denotes the high-frequency shear modulus of the material evaluated at a temperature equal to the glass transition temperature, is the so-called shoving volume, i.e. it is the characteristic volume of the group of atoms involved in the shoving event by which an atom/molecule escapes from the cage of nearest-neighbours, typically on the order of the volume occupied by few atoms. Furthermore, is the thermal expansion coefficient of the material, is a parameter which measures the steepness of the power-law rise of the ascending flank of the first peak of the radial distribution function, and is quantitatively related to the repulsive part of the interatomic potential. Finally, denotes the Boltzmann constant.

Eddy viscosity

In the study of turbulence in fluids, a common practical strategy is to ignore the small-scale vortices 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. 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.

Selected substances

Observed values of viscosity vary over several orders of magnitude, even for common substances. For instance, a 70% sucrose solution has a viscosity over 400 times that of water, and 26000 times that of air. More dramatically, pitch has been estimated to have a viscosity 230 billion times that of water.

Water

The dynamic viscosity of water is about 0.89 mPa·s at room temperature. As a function of temperature in kelvins, the viscosity can be estimated using the semi-empirical Vogel-Fulcher-Tammann equation:
where A = 0.02939 mPa·s, B = 507.88 K, and C = 149.3 K. Experimentally determined values of the viscosity are also given in the table below.
Temperature Viscosity
101.3059
201.0016
300.79722
500.54652
700.40355
900.31417

Air

Under standard atmospheric conditions, 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, one is:
which is accurate in the range -20 °C to 400 °C. For this formula to be valid, the temperature must be given in kelvins; then corresponds to the viscosity in Pa·s.
being drizzled

Other common substances

SubstanceViscosity Temperature Ref.
Benzene0.60425
Water1.001620
Mercury1.52625
Whole milk2.1220
Olive oil56.226
Honey2000–1000020
Ketchup5000–2000025
Peanut butter104–106
Pitch10–30

Order of magnitude estimates

The following table illustrates the range of viscosity values observed in common substances. Unless otherwise noted, a temperature of 25 °C and a pressure of 1 atmosphere are assumed. Certain substances of variable composition or with non-Newtonian behavior are not assigned precise values, since in these cases viscosity depends on additional factors besides temperature and pressure.
Factor DescriptionExamplesValues Ref.
10−6Lower range of gaseous viscosity
Butane7.49 × 10−6
10−6Lower range of gaseous viscosity
Hydrogen8.8 × 10−6
10−5Upper range of gaseous viscosityKrypton2.538 × 10−5
10−5Upper range of gaseous viscosityNeon3.175 × 10−5
10−4Lower range of liquid viscosityPentane2.24 × 10−4
10−4Lower range of liquid viscosityGasoline6 × 10−4
10−4Lower range of liquid viscosityWater8.90 × 10−4
10−3Typical range for small-molecule
Newtonian liquids
Ethanol1.074 × 10−3
10−3Typical range for small-molecule
Newtonian liquids
Mercury1.526 × 10−3
10−3Typical range for small-molecule
Newtonian liquids
Whole milk 2.12 × 10−3
10−3Typical range for small-molecule
Newtonian liquids
Blood4 × 10−3
10−2 – 100Oils and long-chain hydrocarbonsLinseed oil0.028
10−2 – 100Oils and long-chain hydrocarbonsOlive oil0.084
10−2 – 100Oils and long-chain hydrocarbonsSAE 10 Motor oil0.085 to 0.14
10−2 – 100Oils and long-chain hydrocarbonsCastor oil0.1
10−2 – 100Oils and long-chain hydrocarbonsSAE 20 Motor oil0.14 to 0.42
10−2 – 100Oils and long-chain hydrocarbonsSAE 30 Motor oil0.42 to 0.65
10−2 – 100Oils and long-chain hydrocarbonsSAE 40 Motor oil0.65 to 0.90
10−2 – 100Oils and long-chain hydrocarbonsGlycerine1.5
10−2 – 100Oils and long-chain hydrocarbonsPancake syrup2.5
101 – 103Pastes, gels, and other semisolids
Ketchup≈ 101
101 – 103Pastes, gels, and other semisolids
Mustard≈ 101
101 – 103Pastes, gels, and other semisolids
Sour cream≈ 102
101 – 103Pastes, gels, and other semisolids
Peanut butter≈ 102
101 – 103Pastes, gels, and other semisolids
Lard≈ 103
≈108Viscoelastic polymersPitch
≈1021Certain solids under a viscoelastic
description
Mantle

Footnotes

Citations