List of viscosities

Last updated

Dynamic viscosity is a material property which describes the resistance of a fluid to shearing flows. It corresponds roughly to the intuitive notion of a fluid's 'thickness'. For instance, honey has a much higher viscosity than water. Viscosity is measured using a viscometer. Measured values span several orders of magnitude. Of all fluids, gases have the lowest viscosities, and thick liquids have the highest.

Contents

The values listed in this article are representative estimates only, as they do not account for measurement uncertainties, variability in material definitions, or non-Newtonian behavior.

Kinematic viscosity is dynamic viscosity divided by fluid density. This page lists only dynamic viscosity.

Units and conversion factors

For dynamic viscosity, the SI unit is Pascal-second. In engineering, the unit is usually Poise or centiPoise, with 1 Poise = 0.1 Pascal-second, and 1 centiPoise = 0.01 Poise.

For kinematic viscosity, the SI unit is m^2/s. In engineering, the unit is usually Stoke or centiStoke, with 1 Stoke = 0.0001 m^2/s, and 1 centiStoke = 0.01 Stoke.

For liquid, the dynamic viscosity is usually in the range of 0.001 to 1 Pascal-second, or 1 to 1000 centiPoise. The density is usually on the order of 1000 kg/m^3, i.e. that of water. Consequently, if a liquid has dynamic viscosity of n centiPoise, and its density is not too different from that of water, then its kinematic viscosity is around n centiStokes.

For gas, the dynamic viscosity is usually in the range of 10 to 20 microPascal-seconds, or 0.01 to 0.02 centiPoise. The density is usually on the order of 0.5 to 5 kg/m^3. Consequently, its kinematic viscosity is around 2 to 40 centiStokes.

Viscosities at or near standard conditions

Here "standard conditions" refers to temperatures of 25 °C and pressures of 1 atmosphere. Where data points are unavailable for 25 °C or 1 atmosphere, values are given at a nearby temperature/pressure.

The temperatures corresponding to each data point are stated explicitly. By contrast, pressure is omitted since gaseous viscosity depends only weakly on it.

Gases

Noble gases

The simple structure of noble gas molecules makes them amenable to accurate theoretical treatment. For this reason, measured viscosities of the noble gases serve as important tests of the kinetic-molecular theory of transport processes in gases (see Chapman–Enskog theory). One of the key predictions of the theory is the following relationship between viscosity , thermal conductivity , and specific heat :

where is a constant which in general depends on the details of intermolecular interactions, but for spherically symmetric molecules is very close to . [1]

This prediction is reasonably well-verified by experiments, as the following table shows. Indeed, the relation provides a viable means for obtaining thermal conductivities of gases since these are more difficult to measure directly than viscosity. [1] [2]

SubstanceMolecular
formula
Viscosity
(μPa·s)
Thermal conductivity
(W m−1K−1)
Specific heat
(J K−1kg−1)
NotesRefs.
Helium He19.850.15331162.47 [2] [3]
Neon Ne31.750.04926182.51 [2] [3]
Argon Ar22.610.01783132.52 [2] [3]
Krypton Kr25.380.00941492.49 [2] [3]
Xenon Xe23.080.005695.02.55 [2] [3]
Radon Rn≈26≈0.0036456.2T = 26.85 °C;
calculated theoretically;
estimated assuming
[4]

Diatomic elements

SubstanceMolecular formulaViscosity (μPa·s)NotesRef.
Hydrogen H28.90 [5]
Nitrogen N217.76 [5]
Oxygen O220.64 [6]
Fluorine F223.16 [7]
Chlorine Cl213.40 [7]

Hydrocarbons

SubstanceMolecular formulaViscosity (μPa·s)NotesRef.
Methane CH411.13 [8]
Acetylene C2H210.2T = 20 °C [9]
Ethylene C2H410.28 [8]
Ethane C2H69.27 [8]
Propyne C3H48.67T = 20 °C [9]
Propene C3H68.39 [10]
Propane C3H88.18 [8]
Butane C4H107.49 [8]

Organohalides

SubstanceMolecular formulaViscosity (μPa·s)NotesRef.
Carbon tetrafluoride CF417.32 [11]
Fluoromethane CH3F11.79 [12]
Difluoromethane CH2F212.36 [12]
Fluoroform CHF314.62 [12]
Pentafluoroethane C2HF512.94 [12]
Hexafluoroethane C2F614.00 [12]
Octafluoropropane C3F812.44 [12]

Other gases

SubstanceMolecular formulaViscosity (μPa·s)NotesRef.
Air 18.46 [6]
Ammonia NH310.07 [13]
Nitrogen trifluoride NF317.11T = 26.85 °C [14]
Boron trichloride BCl312.3Theoretical estimate at T = 26.85 °C;
estimated uncertainty of 10%
[14]
Carbon dioxide CO214.90 [15]
Carbon monoxide CO17.79 [16]
Hydrogen sulfide H2S12.34 [17]
Nitric oxide NO18.90 [7]
Nitrous oxide N2O14.90 [18]
Sulfur dioxide SO212.82 [10]
Sulfur hexafluoride SF615.23 [5]
Molybdenum hexafluoride MoF614.5Theoretical estimates at T = 26.85 °C [19]
Tungsten hexafluoride WF617.1
Uranium hexafluoride UF617.4

Liquids

n-Alkanes

Substances composed of longer molecules tend to have larger viscosities due to the increased contact of molecules across layers of flow. [20] This effect can be observed for the n-alkanes and 1-chloroalkanes tabulated below. More dramatically, a long-chain hydrocarbon like squalene (C30H62) has a viscosity an order of magnitude larger than the shorter n-alkanes (roughly 31 mPa·s at 25 °C). This is also the reason oils tend to be highly viscous, since they are usually composed of long-chain hydrocarbons.

SubstanceMolecular formulaViscosity (mPa·s)NotesRef.
Pentane C5H120.224 [21]
Hexane C6H140.295 [22]
Heptane C7H160.389 [22]
Octane C8H180.509 [22]
Nonane C9H200.665 [21]
Decane C10H220.850 [22]
Undecane C11H241.098 [21]
Dodecane C12H261.359 [22]
Tridecane C13H281.724 [21]
Tetradecane C14H302.078 [22]
Pentadecane C15H322.82T = 20 °C [23]
Hexadecane C16H343.03 [21]
Heptadecane C17H364.21T = 20 °C [24]

1-Chloroalkanes

SubstanceMolecular formulaViscosity (mPa·s)NotesRef.
Chlorobutane C4H9Cl0.4261 [25]
Chlorohexane C6H11Cl0.6945
Chlorooctane C8H17Cl1.128
Chlorodecane C10H21Cl1.772
Chlorododecane C12H25Cl2.668
Chlorotetradecane C14H29Cl3.875
Chlorohexadecane C16H33Cl5.421
Chlorooctadecane C18H37Cl7.385Supercooled liquid

Other halocarbons

SubstanceMolecular formulaViscosity (mPa·s)NotesRef.
Dichloromethane CH2Cl20.401 [26]
Trichloromethane
(chloroform)
CHCl30.52 [10]
Tribromomethane
(bromoform)
CHBr31.89 [27]
Carbon tetrachloride CCl40.86 [27]
Trichloroethylene C2HCl30.532 [28]
Tetrachloroethylene C2Cl40.798T = 30 °C [28]
Chlorobenzene C6H5Cl0.773 [29]
Bromobenzene C6H5Br1.080 [29]
1-Bromodecane C10H21Br3.373 [30]

Alkenes

SubstanceMolecular formulaViscosity (mPa·s)NotesRef.
2-Pentene C5H100.201 [31]
1-Hexene C6H120.271 [32]
1-Heptene C7H140.362 [32]
1-Octene C8H160.506T = 20 °C [31]
2-Octene C8H160.506T = 20 °C [31]
n-Decene C10H200.828T = 20 °C [31]

Other liquids

SubstanceMolecular formulaViscosity (mPa·s)NotesRef.
Acetic acid C2H4O21.056 [21]
Acetone C3H6O0.302 [33]
Benzene C6H60.604 [21]
Bromine Br20.944 [21]
Ethanol C2H6O1.074 [21]
Glycerol C3H8O31412 [34]
Hydrazine H4N20.876 [21]
Iodine pentafluoride IF52.111 [35]
Mercury Hg1.526 [21]
Methanol CH4O0.553 [36]
1-Propanol (propyl alcohol)C3H8O1.945 [37]
2-Propanol (isopropyl alcohol)C3H8O2.052 [37]
Squalane C30H6231.123 [38]
WaterH2O1.0016T = 20 °C, standard pressure [21]

Aqueous solutions

The viscosity of an aqueous solution can either increase or decrease with concentration depending on the solute and the range of concentration. For instance, the table below shows that viscosity increases monotonically with concentration for sodium chloride and calcium chloride, but decreases for potassium iodide and cesium chloride (the latter up to 30% mass percentage, after which viscosity increases).

The increase in viscosity for sucrose solutions is particularly dramatic, and explains in part the common experience of sugar water being "sticky".

Table: Viscosities (in mPa·s) of aqueous solutions at T = 20 °C for various solutes and mass percentages [21]
Solutemass percentage = 1%2%3%4%5%10%15%20%30%40%50%60%70%
Sodium chloride (NaCl)1.0201.0361.0521.0681.0851.1931.3521.557
Calcium chloride (CaCl2)1.0281.0501.0781.1101.1431.3191.5641.9303.4678.997
Potassium iodide (KI)0.9970.9910.9860.9810.9760.9460.9250.9100.8920.897
Cesium chloride (CsCl)0.9970.9920.9880.9840.9800.9660.9530.9390.9220.9340.9811.120
Sucrose (C12H22O11)1.0281.0551.0841.1141.1461.3361.5921.9453.1876.16215.43158.487481.561

Substances of variable composition

SubstanceViscosity (mPa·s)Temperature (°C)Reference
Whole milk 2.1220 [39]
Blood 2 - 937 [40]
Olive oil 56.226 [39]
Canola oil 46.230 [39]
Sunflower oil 48.826 [39]
Honey 2000-10,00020 [41]
Ketchup [lower-alpha 1] 5000-20,00025 [42]
Peanut butter [lower-alpha 1] 104-106 [43]
Pitch 2.3×101110-30 (variable) [44]
  1. 1 2 These materials are highly non-Newtonian.

Viscosities under nonstandard conditions

Gases

Pressure dependence of the viscosity of dry air at 300, 400 and 500 kelvins Air dry dynamic visocity on pressure temperature.svg
Pressure dependence of the viscosity of dry air at 300, 400 and 500 kelvins

All values are given at 1 bar (approximately equal to atmospheric pressure).

SubstanceChemical formulaTemperature (K)Viscosity (μPa·s)
Air 100 7.1
200 13.3
300 18.5
400 23.1
500 27.1
600 30.8
Ammonia NH3300 10.2
400 14.0
500 17.9
600 21.7
Carbon dioxide CO2200 10.1
300 15.0
400 19.7
500 24.0
600 28.0
Helium He 100 9.6
200 15.1
300 19.9
400 24.3
500 28.3
600 32.2
Water vapor H2O 380 12.498
400 13.278
450 15.267
500 17.299
550 19.356
600 21.425
650 23.496
700 25.562
750 27.617
800 29.657
900 33.680
1000 37.615
1100 41.453
1200 45.192

Liquids (including liquid metals)

Viscosity of water as a function of temperature Dynamic Viscosity of Water.png
Viscosity of water as a function of temperature
SubstanceChemical formulaTemperature (°C)Viscosity (mPa·s)
Mercury [45] [46] Hg -30 1.958
-20 1.856
-10 1.766
0 1.686
10 1.615
20 1.552
25 1.526
30 1.495
50 1.402
75 1.312
100 1.245
126.85 1.187
226.85 1.020
326.85 0.921
Ethanol C2H6O -25 3.26
0 1.786
25 1.074
50 0.694
75 0.476
Bromine Br20 1.252
25 0.944
50 0.746
Water H2O 0.01 1.7911
10 1.3059
20 1.0016
25 0.89002
30 0.79722
40 0.65273
50 0.54652
60 0.46603
70 0.40355
80 0.35405
90 0.31417
99.606 0.28275
Glycerol C3H8O325 934
50 152
75 39.8
100 14.76
Aluminum Al 700 1.24
800 1.04
900 0.90
Gold Au 1100 5.130
1200 4.640
1300 4.240
Copper Cu 1100 3.92
1200 3.34
1300 2.91
1400 2.58
1500 2.31
1600 2.10
1700 1.92
Silver Ag 1300 3.75
1400 3.27
1500 2.91
Iron Fe 1600 5.22
1700 4.41
1800 3.79
1900 3.31
2000 2.92
2100 2.60

In the following table, the temperature is given in kelvins.

SubstanceChemical formulaTemperature (K)Viscosity (mPa·s)
Gallium [46] Ga 400 1.158
500 0.915
600 0.783
700 0.700
800 0.643
Zinc [46] Zn 700 3.737
800 2.883
900 2.356
1000 2.005
1100 1.756
Cadmium [46] Cd 600 2.708
700 2.043
800 1.654
900 1.403

Solids

SubstanceViscosity (Pa·s)Temperature (°C)
granite [47] 3×1019 - 6×101925
asthenosphere [48] 7.0×1019900
upper mantle [48] 7×10201×10211300–3000
lower mantle [ citation needed ]1×10212×10213000–4000

Related Research Articles

<span class="mw-page-title-main">Equation of state</span> An equation describing the state of matter under a given set of physical conditions

In physics and chemistry, an equation of state is a thermodynamic equation relating state variables, which describe the state of matter under a given set of physical conditions, such as pressure, volume, temperature, or internal energy. Most modern equations of state are formulated in the Helmholtz free energy. Equations of state are useful in describing the properties of pure substances and mixtures in liquids, gases, and solid states as well as the state of matter in the interior of stars.

<span class="mw-page-title-main">Lennard-Jones potential</span> Model of intermolecular interactions

In computational chemistry, the Lennard-Jones potential is an intermolecular pair potential. Out of all the intermolecular potentials, the Lennard-Jones potential is probably the one that has been the most extensively studied. It is considered an archetype model for simple yet realistic intermolecular interactions.

<span class="mw-page-title-main">Freezing-point depression</span> Process in which adding a solute to a solvent decreases the freezing point of the solvent

Freezing-point depression is a drop in the maximum temperature at which a substance freezes, caused when a smaller amount of another, non-volatile substance is added. Examples include adding salt into water, alcohol in water, ethylene or propylene glycol in water, adding copper to molten silver, or the mixing of two solids such as impurities into a finely powdered drug.

<span class="mw-page-title-main">Bromobenzene</span> Chemical compound

Bromobenzene is an aryl halide, C6H5Br. It is a colourless liquid although older samples can appear yellow. It is a reagent in organic synthesis.

In thermodynamics, an activity coefficient is a factor used to account for deviation of a mixture of chemical substances from ideal behaviour. In an ideal mixture, the microscopic interactions between each pair of chemical species are the same and, as a result, properties of the mixtures can be expressed directly in terms of simple concentrations or partial pressures of the substances present e.g. Raoult's law. Deviations from ideality are accommodated by modifying the concentration by an activity coefficient. Analogously, expressions involving gases can be adjusted for non-ideality by scaling partial pressures by a fugacity coefficient.

The Benedict–Webb–Rubin equation (BWR), named after Manson Benedict, G. B. Webb, and L. C. Rubin, is an equation of state used in fluid dynamics. Working at the research laboratory of the M. W. Kellogg Company, the three researchers rearranged the Beattie–Bridgeman equation of state and increased the number of experimentally determined constants to eight.

Jürgen Gmehling is a retired German professor of technical and industrial chemistry at the Carl von Ossietzky University of Oldenburg.

<span class="mw-page-title-main">Diphenyl ether</span> Chemical compound

Diphenyl ether is the organic compound with the formula (C6H5)2O. It is a colorless, low-melting solid. This, the simplest diaryl ether, has a variety of niche applications.

Binary liquid is a type of chemical combination, which creates a special reaction or feature as a result of mixing two liquid chemicals, that are normally inert or have no function by themselves. A number of chemical products are produced as a result of mixing two chemicals as a binary liquid, such as plastic foams and some explosives.

<span class="mw-page-title-main">Viscosity</span> Resistance of a fluid to shear deformation

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 is defined scientifically as a force multiplied by a time divided by an area. Thus its SI units are newton-seconds per square metre, or pascal-seconds.

<span class="mw-page-title-main">Liquid</span> State of matter

A liquid is a nearly incompressible fluid that conforms to the shape of its container but retains a nearly constant volume independent of pressure. It is one of the four fundamental states of matter, and is the only state with a definite volume but no fixed shape.

<span class="mw-page-title-main">Cerium(III) bromide</span> Chemical compound

Cerium(III) bromide is an inorganic compound with the formula CeBr3. This white hygroscopic solid is of interest as a component of scintillation counters.

An osmotic coefficient is a quantity which characterises the deviation of a solvent from ideal behaviour, referenced to Raoult's law. It can be also applied to solutes. Its definition depends on the ways of expressing chemical composition of mixtures.

The upper critical solution temperature (UCST) or upper consolute temperature is the critical temperature above which the components of a mixture are miscible in all proportions. The word upper indicates that the UCST is an upper bound to a temperature range of partial miscibility, or miscibility for certain compositions only. For example, hexane-nitrobenzene mixtures have a UCST of 19 °C (66 °F), so that these two substances are miscible in all proportions above 19 °C (66 °F) but not at lower temperatures. Examples at higher temperatures are the aniline-water system at 168 °C (334 °F), and the lead-zinc system at 798 °C (1,468 °F).

Squalane is the organic compound with the formula ( 2CH 3CH 33CH 2)2. A colorless hydrocarbon, it is the hydrogenated derivative of squalene, although commercial samples are derived from nature. In contrast to squalene, due to the complete saturation of squalane, it is not subject to auto-oxidation. This fact, coupled with its lower costs and desirable physical properties, led to its use as an emollient and moisturizer in cosmetics.

<span class="mw-page-title-main">2,3-Dimethylpentane</span> Chemical compound

2,3-Dimethylpentane is an organic compound of carbon and hydrogen with formula C
7
H
16
, more precisely CH
3
CH(CH
3
)
CH(CH
3
)
CH
2
CH
3
: a molecule of pentane with methyl groups –CH
3
replacing hydrogen atoms on carbon atoms 2 and 3. It is an alkane, a fully saturated hydrocarbon; specifically, one of the isomers of heptane.

<i>N</i>,<i>N</i>-diethylmethylamine Organic compound, industrial chemical

N,N-diethylmethylamine (diethylmethylamine, DEMA) is a tertiary amine with the formula C5H13N. N,N-Diethylmethylamine is a clear, colorless to pale yellow liquid at room temperature, and is used in various industrial and scientific applications including water desalination as well as analytical and organic chemistry.

A protic ionic liquid is an ionic liquid that is formed via proton transfer from a Brønsted acid to a Brønsted base. Unlike many other types of ionic liquids, which are formed through a series of synthesis steps, protic ionic liquids are easier to create because the acid and base must simply be mixed together.

Deresh RamjugernathFAAS is a South African professor of Engineering Technology & Applied Sciences. He was a Deputy Vice-Chancellor of Research at the University of KwaZulu-Natal (UKZN) and the current Deputy Vice-Chancellor of Learning and Teaching at Stellenbosch University (SU).

Thermodynamic modelling is a set of different strategies that are used by engineers and scientists to develop models capable of evaluating different thermodynamic properties of a system. At each thermodynamic equilibrium state of a system, the thermodynamic properties of the system are specified. Generally, thermodynamic models are mathematical relations that relate different state properties to each other in order to eliminate the need of measuring all the properties of the system in different states.

References

  1. 1 2 Chapman, Sydney; Cowling, T.G. (1970), The Mathematical Theory of Non-Uniform Gases (3rd ed.), Cambridge University Press
  2. 1 2 3 4 5 6 Kestin, J.; Ro, S. T.; Wakeham, W. A. (1972). "Viscosity of the Noble Gases in the Temperature Range 25–700°C". The Journal of Chemical Physics. 56 (8): 4119–4124. doi: 10.1063/1.1677824 . ISSN   0021-9606.
  3. 1 2 3 4 5 Le Neindre, B.; Garrabos, Y.; Tufeu, R. (1989). "Thermal conductivity of dense noble gases". Physica A: Statistical Mechanics and Its Applications. 156 (1): 512–521. doi:10.1016/0378-4371(89)90137-4. ISSN   0378-4371.
  4. Ho, C. Y.; Powell, R. W.; Liley, P. E. (1972). "Thermal Conductivity of the Elements". Journal of Physical and Chemical Reference Data. 1 (2): 279–421. doi:10.1063/1.3253100. ISSN   0047-2689.
  5. 1 2 3 Assael, M. J.; Kalyva, A. E.; Monogenidou, S. A.; Huber, M. L.; Perkins, R. A.; Friend, D. G.; May, E. F. (2018). "Reference Values and Reference Correlations for the Thermal Conductivity and Viscosity of Fluids". Journal of Physical and Chemical Reference Data. 47 (2): 021501. doi:10.1063/1.5036625. ISSN   0047-2689. PMC   6463310 . PMID   30996494.
  6. 1 2 Kestin, J.; Leidenfrost, W. (1959). "An absolute determination of the viscosity of eleven gases over a range of pressures". Physica. 25 (7–12): 1033–1062. doi:10.1016/0031-8914(59)90024-2. ISSN   0031-8914.
  7. 1 2 3 Yaws, Carl L. (1997), Handbook Of Viscosity: Volume 4: Inorganic Compounds And Elements, Gulf Professional Publishing, ISBN   978-0123958501
  8. 1 2 3 4 5 Kestin, J; Khalifa, H.E.; Wakeham, W.A. (1977). "The viscosity of five gaseous hydrocarbons". The Journal of Chemical Physics. 66 (3): 1132–1134. Bibcode:1977JChPh..66.1132K. doi:10.1063/1.434048.
  9. 1 2 Titani, Toshizo (1930). "The viscosity of vapours of organic compounds. Part II". Bulletin of the Chemical Society of Japan. 5 (3): 98–108. doi: 10.1246/bcsj.5.98 .
  10. 1 2 3 Miller, J.W. Jr.; Shah, P.N.; Yaws, C.L. (1976). "Correlation constants for chemical compounds". Chemical Engineering. 83 (25): 153–180. ISSN   0009-2460.
  11. Kestin, J.; Ro, S.T.; Wakeham, W.A. (1971). "Reference values of the viscosity of twelve gases at 25°C". Transactions of the Faraday Society. 67: 2308–2313. doi:10.1039/TF9716702308.
  12. 1 2 3 4 5 6 Dunlop, Peter J. (1994). "Viscosities of a series of gaseous fluorocarbons at 25 °C". The Journal of Chemical Physics. 100 (4): 3149–3151. doi:10.1063/1.466405.
  13. Iwasaki, Hiroji; Takahashi, Mitsuo (1968). "Studies on the transport properties of fluids at high pressure". The Review of Physical Chemistry of Japan. 38 (1).
  14. 1 2 "Database of the Thermophysical Properties of Gases Used in the Semiconductor Industry | NIST". Archived from the original on 2019-07-11. Retrieved 2019-07-11.
  15. Schäfer, Michael; Richter, Markus; Span, Roland (2015). "Measurements of the viscosity of carbon dioxide at temperatures from (253.15 to 473.15)K with pressures up to 1.2MPa". The Journal of Chemical Thermodynamics. 89: 7–15. doi:10.1016/j.jct.2015.04.015. ISSN   0021-9614.
  16. Kestin, J.; Ro, S. T.; Wakeham, W. A. (1982). "The Viscosity of Carbon-Monoxide and its Mixtures with Other Gases in the Temperature Range 25 - 200°C". Berichte der Bunsengesellschaft für physikalische Chemie. 86 (8): 753–760. doi:10.1002/bbpc.19820860816. ISSN   0005-9021.
  17. Pal, Arun K.; Bhattacharyya, P. K. (1969). "Viscosity of Binary Polar‐Gas Mixtures". The Journal of Chemical Physics. 51 (2): 828–831. doi:10.1063/1.1672075. ISSN   0021-9606.
  18. Takahashi, Mitsuo; Shibasaki-Kitakawa, Naomi; Yokoyama, Chiaki; Takahashi, Shinji (1996). "Viscosity of Gaseous Nitrous Oxide from 298.15 K to 398.15 K at Pressures up to 25 MPa". Journal of Chemical & Engineering Data. 41 (6): 1495–1498. doi:10.1021/je960060d. ISSN   0021-9568.
  19. Zarkova, L.; Hohm, U. (2002). "pVT–Second Virial Coefficients B(T), Viscosity eta(T), and Self-Diffusion rhoD(T) of the Gases: BF3, CF4, SiF4, CCl4, SiCl4, SF6, MoF6, WF6, UF6, C(CH3)4, and Si(CH3)4 Determined by Means of an Isotropic Temperature-Dependent Potential". Journal of Physical and Chemical Reference Data. 31 (1): 183–216. doi:10.1063/1.1433462. ISSN   0047-2689.
  20. chem.libretexts.org (11 March 2016). "Intermolecular Forces in Action: Surface Tension, Viscosity, and Capillary Action". chem.libretexts.org. Archived from the original on 2019-07-24. Retrieved 2019-07-24.
  21. 1 2 3 4 5 6 7 8 9 10 11 12 13 CRC Handbook of Chemistry and Physics, 99th Edition (Internet Version 2018), John R. Rumble, ed., CRC Press/Taylor & Francis, Boca Raton, FL.
  22. 1 2 3 4 5 6 Dymond, J. H.; Oye, H. A. (1994). "Viscosity of Selected Liquid n‐Alkanes". Journal of Physical and Chemical Reference Data. 23 (1): 41–53. doi:10.1063/1.555943. ISSN   0047-2689.
  23. Wu, Jianging; Nhaesi, Abdulghanni H.; Asfour, Abdul-Fattah A. (1999). "Viscosities of Eight Binary Liquidn-Alkane Systems at 293.15 K and 298.15 K". Journal of Chemical & Engineering Data. 44 (5): 990–993. doi:10.1021/je980291f. ISSN   0021-9568.
  24. Doolittle, Arthur K. (1951). "Studies in Newtonian Flow. II. The Dependence of the Viscosity of Liquids on Free‐Space". Journal of Applied Physics. 22 (12): 1471–1475. Bibcode:1951JAP....22.1471D. doi:10.1063/1.1699894. ISSN   0021-8979.
  25. Coursey, B. M.; Heric, E. L. (1971). "AApplication of the Congruence Principle to Viscosities of 1-Chloroalkane Binary Mixtures". Canadian Journal of Chemistry. 49 (16): 2631–2635. doi: 10.1139/v71-437 . ISSN   0008-4042.
  26. Wang, Jianji; Tian, Yong; Zhao, Yang; Zhuo, Kelei (2003). "A volumetric and viscosity study for the mixtures of 1-n-butyl-3-methylimidazolium tetrafluoroborate ionic liquid with acetonitrile, dichloromethane, 2-butanone and N, N ? dimethylformamide". Green Chemistry. 5 (5): 618. doi:10.1039/b303735e. ISSN   1463-9262.
  27. 1 2 Reid, Robert C.; Prausnitz, John M.; Poling, Bruce E. (1987), The Properties of Gases and Liquids, McGraw-Hill Book Company, p. 442, ISBN   0-07-051799-1
  28. 1 2 Venkatesulu, D.; Venkatesu, P.; Rao, M. V. Prabhakara (1997). "Viscosities and Densities of Trichloroethylene or Tetrachloroethylene with 2-Alkoxyethanols at 303.15 K and 313.15 K". Journal of Chemical & Engineering Data. 42 (2): 365–367. doi:10.1021/je960316f. ISSN   0021-9568.
  29. 1 2 Nayak, Jyoti N.; Aralaguppi, Mrityunjaya I.; Aminabhavi, Tejraj M. (2003). "Density, Viscosity, Refractive Index, and Speed of Sound in the Binary Mixtures of Ethyl Chloroacetate + Cyclohexanone, + Chlorobenzene, + Bromobenzene, or + Benzyl Alcohol at (298.15, 303.15, and 308.15) K". Journal of Chemical & Engineering Data. 48 (3): 628–631. doi:10.1021/je0201828. ISSN   0021-9568.
  30. Cokelet, Giles R.; Hollander, Frederick J.; Smith, Joseph H. (1969). "Density and viscosity of mixtures of 1,1,2,2-tetrabromoethane and 1-bromododecane". Journal of Chemical & Engineering Data. 14 (4): 470–473. doi:10.1021/je60043a017. ISSN   0021-9568.
  31. 1 2 3 4 Wright, Franklin J. (1961). "Influence of Temperature on Viscosity of Nonassociated Liquids". Journal of Chemical & Engineering Data. 6 (3): 454–456. doi:10.1021/je00103a035. ISSN   0021-9568.
  32. 1 2 Sagdeev, D. I.; Fomina, M. G.; Mukhamedzyanov, G. Kh.; Abdulagatov, I. M. (2014). "Experimental Study and Correlation Models of the Density and Viscosity of 1-Hexene and 1-Heptene at Temperatures from (298 to 473) K and Pressures up to 245 MPa". Journal of Chemical & Engineering Data. 59 (4): 1105–1119. doi:10.1021/je401015e. ISSN   0021-9568.
  33. Petrino, P. J.; Gaston-Bonhomme, Y. H.; Chevalier, J. L. E. (1995). "Viscosity and Density of Binary Liquid Mixtures of Hydrocarbons, Esters, Ketones, and Normal Chloroalkanes". Journal of Chemical & Engineering Data. 40 (1): 136–140. doi:10.1021/je00017a031. ISSN   0021-9568.
  34. Segur, J. B.; Oberstar, H. E. (1951). "Viscosity of Glycerol and Its Aqueous Solutions". Industrial & Engineering Chemistry. 43 (9): 2117–2120. doi:10.1021/ie50501a040.
  35. Hetherington, G.; Robinson, P.L. (1956). "The Viscosities of Iodine Pentafluoride and Ditellurium Decafluoride". Journal of the Chemical Society (Resumed): 3681. doi:10.1039/jr9560003674. ISSN   0368-1769.
  36. Canosa, J.; Rodríguez, A.; Tojo, J. (1998). "Dynamic Viscosities of (Methyl Acetate or Methanol) with (Ethanol, 1-Propanol, 2-Propanol, 1-Butanol, and 2-Butanol) at 298.15 K". Journal of Chemical & Engineering Data. 43 (3): 417–421. doi:10.1021/je9702302. ISSN   0021-9568.
  37. 1 2 Paez, Susana; Contreras, Martin (1989). "Densities and viscosities of binary mixtures of 1-propanol and 2-propanol with acetonitrile". Journal of Chemical & Engineering Data. 34 (4): 455–459. doi:10.1021/je00058a025. ISSN   0021-9568.
  38. Lal, Krishan; Tripathi, Neelima; Dubey, Gyan P. (2000). "Densities, Viscosities, and Refractive Indices of Binary Liquid Mixtures of Hexane, Decane, Hexadecane, and Squalane with Benzene at 298.15 K". Journal of Chemical & Engineering Data. 45 (5): 961–964. doi:10.1021/je000103x. ISSN   0021-9568.
  39. 1 2 3 4 Fellows, P.J. (2009), Food Processing Technology: Principles and Practice (3rd ed.), Woodhead Publishing, ISBN   978-1845692162
  40. Pries, A. R.; Neuhaus, D.; Gaehtgens, P. (1992-12-01). "Blood viscosity in tube flow: dependence on diameter and hematocrit". American Journal of Physiology. Heart and Circulatory Physiology. 263 (6): H1770–H1778. doi:10.1152/ajpheart.1992.263.6.H1770. ISSN   0363-6135.
  41. Yanniotis, S.; Skaltsi, S.; Karaburnioti, S. (February 2006). "Effect of moisture content on the viscosity of honey at different temperatures". Journal of Food Engineering. 72 (4): 372–377. doi:10.1016/j.jfoodeng.2004.12.017.
  42. Koocheki, Arash; Ghandi, Amir; Razavi, Seyed M. A.; Mortazavi, Seyed Ali; Vasiljevic, Todor (2009), "The rheological properties of ketchup as a function of different hydrocolloids and temperature", International Journal of Food Science & Technology, 44 (3): 596–602, doi:10.1111/j.1365-2621.2008.01868.x
  43. Citerne, Guillaume P.; Carreau, Pierre J.; Moan, Michel (2001), "Rheological properties of peanut butter", Rheologica Acta, 40 (1): 86–96, doi:10.1007/s003970000120, S2CID   94555820
  44. Edgeworth, R; Dalton, B J; Parnell, T (1984), "The pitch drop experiment", European Journal of Physics, 5 (4): 198–200, Bibcode:1984EJPh....5..198E, doi:10.1088/0143-0807/5/4/003, S2CID   250769509
  45. Suhrmann, Von R.; Winter, E.-O. (1955), "Dichte- und Viskositätsmessungen an Quecksilber und hochverdünnten Kalium- und Cäsiumamalgamen vom Erstarrungspunkt bis + 30 C", Zeitschrift für Naturforschung, 10a (12): 985, doi: 10.1515/zna-1955-1211 , S2CID   97692836, archived from the original on 2020-02-15, retrieved 2021-10-17
  46. 1 2 3 4 Assael, Marc J.; Armyra, Ivi J.; Brillo, Juergen; Stankus, Sergei V.; Wu, Jiangtao; Wakeham, William A. (2012), "Reference Data for the Density and Viscosity of Liquid Cadmium, Cobalt, Gallium, Indium, Mercury, Silicon, Thallium, and Zinc" (PDF), Journal of Physical and Chemical Reference Data, 41 (3): 033101, doi:10.1063/1.4729873, archived (PDF) from the original on 2021-10-17, retrieved 2019-12-12
  47. Kumagai, Naoichi; Sasajima, Sadao; Ito, Hidebumi (15 February 1978). "Long-term Creep of Rocks: Results with Large Specimens Obtained in about 20 Years and Those with Small Specimens in about 3 Years". Journal of the Society of Materials Science (Japan). 27 (293): 157–161. Archived from the original on 2011-05-21. Retrieved 2008-06-16.
  48. 1 2 Fjeldskaar, W. (1994). "Viscosity and thickness of the asthenosphere detected from the Fennoscandian uplift". Earth and Planetary Science Letters. 126 (4): 399–410. Bibcode:1994E&PSL.126..399F. doi:10.1016/0012-821X(94)90120-1.