Open access peer-reviewed chapter

Viscosity Models Based on the Free Volume and Entropy Scaling Theories for Pure Hydrocarbons over a Wide Range of Temperatures and Pressures

Written By

Hseen O. Baled and Isaac K. Gamwo

Submitted: 28 March 2019 Reviewed: 12 April 2019 Published: 07 October 2019

DOI: 10.5772/intechopen.86321

From the Edited Volume

Thermophysical Properties of Complex Materials

Edited by Aamir Shahzad

Chapter metrics overview

901 Chapter Downloads

View Full Metrics

Abstract

Viscosity is a critical fundamental property required in many applications in the chemical and oil industry. Direct measurements of this property are usually expensive and time-consuming. Therefore, reliable predictive methods are often employed to obtain the viscosity. In this work, two viscosity models based on the free-volume and entropy scaling theories are assessed and compared for pure hydrocarbons. The modeling results are compared to experimental data of 52 pure hydrocarbons including straight-chain alkanes, branched alkanes, cycloalkanes, and aromatics. This study considers viscosity data to extremely high-temperature and high-pressure (HTHP) conditions up to 573 K and 300 MPa. The results obtained with the free-volume theory viscosity in conjunction with the perturbed-chain statistical associating fluid theory (PC-SAFT) equation of state are characterized by an overall average absolute deviation (AAD%) of 3% from the experimental data. The overall AAD% obtained with the predictive entropy scaling method by Lötgering-Lin and Gross is 8%.

Keywords

  • high temperature
  • high pressure
  • hydrocarbons
  • modeling
  • viscosity

1. Introduction

Viscosity is a key property in many engineering disciplines, including chemical and petroleum engineering. For instance, viscosity influences the fluid flow through porous media and pipelines; hence, it is required for the design of pipelines and transport equipment as well as for the estimation of recoverable oil and flow rates in porous media or wellbores. Viscosity can be determined through experimental measurements. However, carrying out viscosity measurements at all conditions of interest is not only expensive and time-consuming but also may not be possible at extreme conditions such as those encountered in ultra-deep reservoirs including pressures up to 300 MPa and temperatures up to 573 K. Reliable prediction models provide an alternative approach to generating predicted and correlated viscosity data at conditions where experimental data are not readily available.

Unlike the viscosity of gases at low pressures which is well defined by the kinetic theory of the gases, the viscosity theory of liquids is still inadequately developed due to the complications caused by the intermolecular forces between the molecules [1]. Therefore, there is no widely accepted simple theoretical method for predicting liquid viscosities, and most estimation techniques used for viscosity prediction of liquids are of empirical or semiempirical nature. The empirical models are correlations based on experimental observation with no theoretical background, whereas semi-theoretical models have a fundamental basis but contain adjustable parameters determined by fitting the model to experimental data.

In the present study, two viscosity models based on the free-volume and entropy scaling theories are assessed and tested against viscosity data for pure hydrocarbons from different chemical families that are commonly found in crude oil at high-temperature and high-pressure (HTHP) conditions up to 573 K and 300 MPa. Pure components are well-suited for initial evaluation of the viscosity models because a viscosity model that has difficulty in correctly describing the viscosity of a single hydrocarbon is likely to fail when predicting multicomponent mixtures.

Advertisement

2. Free-volume theory (FVT)

The FVT model is based on the free-volume concept. The idea that the viscosity depends upon the free space was first introduced by Batschinski [2] about 100 years ago. The viscosity, η, can be expressed as a sum of two contributions given in Eq. (1):

η=η0+ΔηE1

where η0 is the dilute gas viscosity and the Δη term dominates for liquid viscosity. The dilute gas term η0 is determined from the kinetic gas theory at very low pressures. It should be noted, though, that for liquids and supercritical fluids, the dilute gas viscosity term η0 is negligibly small in comparison to the total viscosity η; hence, η0 can be neglected for such fluids. Doolittle [3] found that the viscosity of liquid n-alkanes can be represented by a simple function of the free space fraction, fv=vfv0=vv0v0:

η=AexpB/fvE2

where v0 is the molecular volume of reference or hard-core volume, v is the specific molecular volume, B is characteristic of the free-volume overlap, and A is a material-specific constant. A viscosity model based on the relation between free volume, friction coefficient, and viscosity has been proposed by Allal and coauthors [4, 5]:

η=AexpB/fvE3

where the free-volume fraction fv was defined by means of the fluctuation-dissipation theory as

fvRTE32E4

where R is the gas constant and T is the temperature. In this expression, E=E0+PMρ, where E0=αρ is related to the energy barrier that the molecule has to overcome in order to diffuse, ρ is the density, and P is the pressure. The viscosity of the dense state is linked to the fluid microstructure using the friction coefficient, ζ, which is related to molecular mobility and to the diffusion of linear momentum:

η=ρNAζL2ME5

where NA is Avogadro’s number and L is a characteristic molecular length parameter. By combining the relations between free volume, friction coefficient, and viscosity, the following expression was obtained for the viscosity:

η=ρlαρ+PMρ3RTMexpBαρ+PMρRT3/2E6

The unit for the viscosity is [Pa·s], when all other variables are in SI units. The term PMρ is linked to the energy necessary to form vacant vacuums required for the diffusion of the molecules. l is the characteristic length parameter in Å. The unitless parameter B is characteristic of the free-volume overlap. The density appears explicitly in Eq. (6), and hence the values of the free-volume theory parameters are directly dependent on whether experimental or calculated densities are used.

The three pure component parameters l, α, and B are determined by fitting Eq. (1) to experimental viscosity data. The use of the FVT requires density information, either experimental or calculated values. In this study, FVT is used in conjunction with the hybrid group-contribution perturbed-chain statistical associating fluid theory equation of state (G-C PC-SAFT EoS) [6] since this EoS provides reliable density predictions over wide ranges of pressure and temperature. In this equation, the PC-SAFT parameters are determined using two different sets of group-contribution (G-C) parameters for two pressure ranges: low-to-moderate pressures (≲7 MPa) and high pressures (≳7 MPa).

Advertisement

3. Entropy scaling model by Lötgering-Lin and Gross (ES-LG)

The basic idea of this method is to relate the viscosity to the residual entropy. The residual entropy is defined as the difference between a real state value and ideal gas state value at the same temperature and density, sresρT=sρTsidρT. Lötgering-Lin and Gross [7] proposed a predictive entropy scaling method for viscosities using a group-contribution (G-C) method based on the group-contribution perturbed-chain polar statistical associating fluid theory equation of state (G-C PCP-SAFT EoS) [8, 9]. Lötgering-Lin and Gross linked the Chapman-Enskog viscosity to PCP-SAFT segments in terms of G-C parameters, with

ηCE,gc=516MmolkBT/mgcNAπσgc2Ωgc22E7

where Mmol is the molar mass, kB is the Boltzmann constant, and T is the absolute temperature, NA is Avogadro’s number, m is the segment number, σ is the segment diameter, and Ω22 is the reduced collision integral. The index gc indicates pure component parameters that are calculated with the group-contribution method based on G-C-PCP-SAFT EoS.

A reduced viscosity is then defined as

η=ηηCE,gcE8

where the pure component reduced viscosity, ηi, is empirically correlated as

lnηi=Ai+Biz+Ciz2+Diz3E9

with

z=sreskBmgc,iE10

The residual entropy, sres, is calculated from the G-C-PCP-SAFT EoS originally proposed by Vijande et al. [8] and reparametrized by Sauer et al. [9]:

sresρT=aresTρE11

where ares=Ares/N is the specific Helmholtz energy given by Gross and Sadowski [10]. N is the total number of molecules.

The viscosity parameters Ai to Di of pure substances are obtained from parameters Aα to Dα of functional group α, respectively. The following empirical expressions are proposed by Lötgering-Lin and Gross [7] for mixing group-contribution parameters:

Ai=αnα,imασα3AαE12
Bi=αnα,imασα3Vtot,iγBαE13
Ci=αnα,iCαE14
Di=Dαnα,iE15

With

Vtot,i=αnα,imασα3E16

where nα,i denotes the number of functional groups of type α in the substance i. The exponent γ and the parameter D are kept constant for all studied substances and are optimized for n-alkanes (D = −0.01245 and γ = 0.45) [7]. The group-contribution parameters Aα, Bα, and Cα of all groups α are given in [7].

Advertisement

4. Modeling results

The two viscosity methods, FVT and ES-LG, are tested on a database consisting of 52 hydrocarbons (21 normal alkanes, 13 branched alkanes, 4 cycloalkanes, 14 aromatics) typically present in most of the crude oils from ambient conditions to extremely high-temperature and high-pressure (HTHP) conditions up to 573 K and 300 MPa. The temperature and pressure ranges considered in this study are given for each substance in Table 1. The performance of each model is assessed by the following statistical measures:

CompoundRanges of conditionsReferenceFVTES-LG
T/KP/MPaAAD/%MD/%Bias/%AAD/%MD/%Bias/%
Straight-chain alkanes (normal alkanes)
CH4298–5730.1–300[11, 12, 13]333−2154713
C2H6298–5730.1–70[11, 14]251735−6
C3H8298–5000.1–100[14, 15]5211620−2
n-C4H10298–5730.1–69[11]42025212
n-C5H12298–5730.1–252[16, 17]3181615−3
n-C6H14298–5730.1–300[16, 18]418−2519−1
n-C7H16298–5730.1–100[16, 19]2604174
n-C8H18298–5230.1–242[18, 20, 21]260272
n-C9H20298–4730.1–300[16, 19, 21]2904104
n-C10H22298–5730.1–300[16, 17, 21, 22, 23, 24]2604142
n-C11H24303–3230.1–62[19]0.11014−1
n-C12H26298–5730.1–300[16, 22, 23, 24, 25, 26, 27]31304172
n-C13H28303–3530.1–100[28]14148−4
n-C14H30313–3930.69–60[29]28−1512−2
n-C15H32310–4080.1–320[25]2704110
n-C16H34298–5340.1–273[30, 31]39010241
n-C17H36323–5730.1–0.1[3]26−2816−8
n-C18H38326–5340.1–280[25, 31]31209288
n-C19H40333–523[32]3804102
n-C20H42326–5341.38–243[31, 33]413012409
n-C32H66373–4580.3–0.3[34]120242
Branched alkanes
Isobutane300–5110.1–55[35, 36]3112313−1
Isopentane303–5730.098–196[16, 37]622−31123−7
Neopentane311–4440.7–55[38]723−31660−13
2-Methylpentane298–5500.1–300[16]1905161
3-Methylpentane3130.1–147[39]1418108
2,2-Dimethylbutane3130.1–147[39]28−12737−27
2,3-Dimethylbutane3130.1–147[39]0.310711−7
3-Ethylpentane3130.1–147[39]23−2172017
2,4-Dimethylpentane31301.147[39]261232923
2,2,4-Trimethylpentane298–5230.1–300[20, 40, 41, 42]2805131
2,3,4-Trimethylpentane298–4530.1–195[41]3110715−7
Squalane303–4731–202[43, 44]633−2264923
2,2,4,4,6,8,8-Heptamethylnonane298–4530.1–195[41, 45, 46]749−6751−5
Cycloalkanes
Cyclopentane298–3530.1–300[47, 48]516−226−1
Cyclohexane298–3930.1–100[49, 50]2704141
Methylcyclohexane298–3430.1–300[51, 52]160410−4
Ethylcyclohexane300–5301–50[53]312−1723−5
Aromatics
Benzene298–3730.1–300[54]130513−4
Toluene298–3730.1–299[55]260392
Ethylbenzene298–4530.1–195[41]241164016
Butylbenzene313–3730.1–100[56]2904114
Hexylbenzene313–3730.1–100[56]130595
Octylbenzene313–3730.1–100[56]150493
1,2-Diphenylethane353–4530.1–195[41]1201217−12
m-Xylene298–4730.1–199[20]260410−2
o-Xylene298–3480.1–110[57]1302530−25
p-Xylene298–3480.1–110[57]27−168−6
Naphthalene375–4540.1–101[58]83311122−11
1-Methylnaphthalene298–4730.1–200[20]1135−73260−32
Tetralin298–4480.1–201[20]520−1295328
Phenanthrene396–5730.1–101[58]1341−61729−17
Overall AAD%38

Table 1.

Performance of FVT and ES-LG models for pure hydrocarbons over wide ranges of pressure and temperature (entries are rounded to nearest whole number).

Absolute Average DeviationAAD=100Ni=1Nηi,calηi,expηi,expE17
Maximum DeviationMD=100·maxηi,calηi,expηi,expE18
Bias=100Ni=1Nηi,calηi,expηi,expE19

where N is the total number of data points, ηi,cal represents the calculated viscosity value, and ηi,exp is the experimental data point obtained from the literature. The absolute average deviation AAD is a measure of how close the calculated values are to the experimental data, while the bias indicates how well the calculated values are distributed around the literature data. Low values of the bias imply that the deviations are evenly distributed about zero. A positive bias indicates overestimation of the calculated viscosity, whereas a negative value indicates underestimation. These statistical measures of the ability of each of the selected seven models to reproduce viscosity values at HTHP conditions for each of the 52 pure compounds are given in Table 1.

The overall AAD obtained with the FVT model in conjunction with the hybrid G-C PC-SAFT EoS is 3%. The three adjustable parameters (l, α, B) required in this method are obtained by fitting the FVT predictions for each pure compound to the corresponding literature data. These optimized parameters yield reliable viscosity values over the whole ranges of temperatures to 573 K and pressures to 300 MPa.

The results obtained with the entropy scaling method by Lötgering-Lin and Gross are generally in good agreement with experimental data with an overall AAD of 8%. The AADs obtained for n-alkanes, branched alkanes, cycloalkanes, and aromatics are within 1–15, 3–27, 2–7, and 3–32%, respectively. These results are impressive for a fully predictive model that requires only the input of the molecular mass and the number of functional groups in each molecule. Unfortunately, this model cannot differentiate between isomers, such as 2-methylpentane and 3-methylpentane, and xylene isomers. In addition, this model has not yet been extended to binary, ternary, and multicomponent mixtures, such as crude oils.

For comparison purposes, Figure 1(a–d) shows the performance of the two studied viscosity methods, FVT and ES-LG, for four pure compounds representative of straight-chain alkanes (n-hexane), branched alkanes (2,2,4-trimethylpentane), cycloalkanes (methylcyclohexane), and aromatics (toluene).

Figure 1.

Viscosity predictions obtained with FVT and ES-LG viscosity models compared with literature data (EXP) for (a) n-hexane, (b) 2,2,4-trimethylpentane, (c) methylcyclohexane, and (d) toluene.

Advertisement

5. Conclusions

This work provides an assessment of the capabilities of two viscosity methods based on the free-volume and entropy scaling theories to model the viscosity of pure hydrocarbons over wide ranges of temperatures and pressures. The performance of the two studied viscosity models is discussed and evaluated by comparison to experimental viscosity data of 52 pure hydrocarbons from four different chemical families, namely, straight-chain alkanes, branched alkanes, cycloalkanes, and aromatics, at ambient and extremely high-temperature and high-pressure (HTHP) conditions up to 573 K and 300 MPa. The viscosity of pure components is required in most mixture models as an input parameter, and hence accurate and reliable model for pure compounds, particularly under high-pressure conditions, is a prerequisite for the mixture viscosity to be accurately estimated using the mixture model. The predictive entropy scaling method proposed by Lötgering-Lin and Gross (ES-LG model) predicts the viscosity with an overall absolute average deviation of about 8%, and the predictions are reasonable for most engineering and industrial applications given that the accuracy of most experimental viscosity data is within 1–5%. The free-volume theory (FVT) viscosity model provides very satisfactory results with an overall AAD of 3%. However, it is important to note that unlike the entropy scaling method, the free-volume theory is not a predictive model and requires that sufficient experimental viscosity data are readily available over the temperature and pressure ranges of interest to determine the fluid specific parameters.

Advertisement

Acknowledgments

The Open Access Publishing fee for this chapter was fully paid by the University Library System, University of Pittsburgh, USA.

References

  1. 1. Poling BE, Prausnitz JM, O’Connell JP. The Properties of Gases and Liquids. 5th ed. New York: McGraw-Hill; 2001
  2. 2. Batschinski AJ. Untersuchungen über die innere Reibung der Fluessigkeiten. Zeitschrift für Physikalische Chemie. 1913;84:644-650
  3. 3. Doolittle AK. Studies in Newtonian flow. II. The dependence of the viscosity of liquids on free-space. Journal of Applied Physics. 1951;22:1471-1475
  4. 4. Allal A, Moha-Ouchane M, Boned C. A new free volume model for dynamic viscosity and density of dense fluids versus pressure and temperature. Physics and Chemistry of Liquids. 2001;39:1-30
  5. 5. Allal A, Boned C, Baylaucq A. Free-volume viscosity model for fluids in the dense and gaseous states. Physical Review E. 2001;64:011203/1-011203/10
  6. 6. Burgess WA, Tapriyal D, Gamwo IK, Wu Y, McHugh MA, Enick RM. New group-contribution parameters for the calculation of PC-SAFT parameters for use at pressures to 276 MPa and temperatures to 533 K. Industrial and Engineering Chemistry Research. 2014;53:2520-2528
  7. 7. Lötgering-Lin O, Gross J. Group contribution method for viscosities based on entropy scaling using the perturbed-chain polar statistical associating fluid theory. Industrial and Engineering Chemistry Research. 2015;54:7942-7952
  8. 8. Vijande J, Pineiro MM, Bessieres D, Saint-Guirons H, Legido JL. Description of PVT behaviour of hydrofluoroethers using the PC-SAFT EOS. Physical Chemistry Chemical Physics. 2004;6:766-770
  9. 9. Sauer E, Stavrou M, Gross J. Comparison between a homo-and a heterosegmented group contribution approach based on the perturbed-chain polar statistical associating fluid theory equation of state. Industrial and Engineering Chemistry Research. 2014;53:14854-14864
  10. 10. Gross J, Sadowski G. Perturbed-chain SAFT: An equation of state based on a perturbation theory for chain molecules. Industrial and Engineering Chemistry Research. 2001;40:1244-1260
  11. 11. Younglove BA, Ely JF. Thermophysical properties of fluids. II. Methane, ethane, propane, isobutane, and normal butane. Journal of Physical and Chemical Reference Data. 1987;16:577-798
  12. 12. Evers C, Lösch H, Wagner W. An absolute viscometer-densimeter and measurements of the viscosity of nitrogen, methane, helium, neon, argon, and krypton over a wide range of density and temperature. International Journal of Thermophysics. 2002;23:1411-1439
  13. 13. van der Gulik PS, Mostert R, van den Berg HR. The viscosity of methane at 25°C up to 10 kbar. Physica A: Statistical Mechanics and its Applications. 1988;151:153-166
  14. 14. Seibt D, Vo K, Herrmann S, Vogel E, Hassel E. Simultaneous viscosity-density measurements on ethane and propane over a wide range of temperature and pressure including the near-critical region. Journal of Chemical & Engineering Data. 2011;56:1476-1493
  15. 15. Vogel E, Küchenmeister C, Bich E, Laesecke A. Reference correlation of the viscosity of propane. Journal of Physical and Chemical Reference Data. 1998;27:947-970
  16. 16. Lemmon EW, McLinden MO, Friend DG. Thermophysical properties of fluid systems. In: Linstrom PJ, Mallard WG, editors. NIST Chemistry WebBook, NIST Standard Reference Database Number 69. Gaithersburg, MD: National Institute of Standards and Technology; retrieved 2018
  17. 17. Oliveira C, Wakeham W. The viscosity of five liquid hydrocarbons at pressures up to 250 MPa. International Journal of Thermophysics. 1992;13:773-790
  18. 18. Dymond JH, Robertson J, Isdale JD. Transport properties of nonelectrolyte liquid mixtures-III. Viscosity coefficients for n-octane, n-decane, and equimolar mixtures of n-octane + n-dodecane and n-hexane + n-dodecane from 25 to 100°C at pressures up to the freezing pressure or 500 MPa. International Journal of Thermophysics. 1981;2:133-154
  19. 19. Assael MJ, Papadaki M. Measurements of the viscosity of n-heptane, n-nonane, and n-undecane at pressures up to 70 MPa. International Journal of Thermophysics. 1991;12:801-810
  20. 20. Caudwell DR, Trusler JPM, Vesovic V, Wakeham WA. Viscosity and density of five hydrocarbon liquids at pressures up to 200 MPa and temperatures up to 473 K. Journal of Chemical & Engineering Data. 2009;54:359-366
  21. 21. Huber ML, Laesecke A, Xiang HW. Viscosity correlations for minor constituent fluids in natural gas: n-octane, n-nonane and n-decane. Fluid Phase Equilibria. 2005;228-229:401-408
  22. 22. Knapstad B, Skolsvik PA, Oye HA. Viscosity of pure hydrocarbons. Journal of Chemical & Engineering Data. 1989;34:37-43
  23. 23. Knapstad B, Skjolsvik PA, Oye HA. Viscosity of three binary hydrocarbon mixtures. Journal of Chemical & Engineering Data. 1991;36:84-88
  24. 24. Dymond JH, Young KJ. Transport properties of nonelectrolyte liquid mixtures. I. Viscosity coefficients for n-alkane mixtures at saturation pressure from 283 to 378 K. International Journal of Thermophysics. 1980;1:331-344
  25. 25. Hogenboom DL, Webb W, Dixon JA. Viscosity of several liquid hydrocarbons as a function of temperature, pressure, and free volume. The Journal of Chemical Physics. 1967;46:2586-2598
  26. 26. Huber ML, Laesecke A, Perkins RA. Transport properties of n-dodecane. Energy & Fuels. 2004;18:968-975
  27. 27. Giller EB, Drikamer HG. Viscosity of normal paraffins near the freezing point. Industrial and Engineering Chemistry. 1949;41:2067-2069
  28. 28. Daugé P, Canet X, Baylaucq A, Boned C. Measurements of the density and viscosity of the tridecane + 2,2,4,4,6,8,8-heptamethylnonane mixtures in the temperature range 293.15-353.15 K at pressures up to 100 MPa. High Temperatures - High Pressures. 2001;33:213-230
  29. 29. Hernández-Galván MA, García-Sánchez F, Macías-Salinas R. Liquid viscosities of benzene, n-tetradecane, and benzene + n-tetradecane from 313 to 393 K and pressures up to 60 MPa: Experiment and modeling. Fluid Phase Equilibria. 2007;262:51-60
  30. 30. Dymond JH, Young KJ, Isdale JD. Transport properties of nonelectrolyte liquid mixtures-II. Viscosity coefficients for the n-hexane + n-hexadecane system at temperatures from 25 to 100°C at pressures up to the freezing pressure or 500 MPa. International Journal of Thermophysics. 1980;1(4):345-373
  31. 31. Baled HO, Xing D, Katz H, Tapriyal D, Gamwo IK, Soong Y, et al. Viscosity of n-hexadecane, n-octadecane and n-eicosane at pressures up to 243 MPa and temperatures up to 534 K. The Journal of Chemical Thermodynamics. 2014;72:108-116
  32. 32. Guseinov S, Naziev Y. Izvestiya Vysshikh Uchebnykh Zavedenii, Russian (USSR). Neft i Gaz. 1974;17:38-62
  33. 33. Rodden JB, Erkey C, Akgerman A. High-temperature diffusion, viscosity, and density measurements in n-eicosane. Journal of Chemical & Engineering Data. 1988;33:344-347
  34. 34. Aasen E, Rytter E, Oye HA. Viscosity of n-hydrocarbons and their mixtures. Industrial and Engineering Chemistry Research. 1990;29:1635-1640
  35. 35. Diller DE, Van Poolen LJ. Measurements of the viscosities of saturated and compressed liquid normal butane and isobutane. International Journal of Thermophysics. 1985;6:43-62
  36. 36. Gonzalez MH, Lee AL. Viscosity of isobutene. Journal of Chemical & Engineering Data. 1966;11:357-359
  37. 37. Bridgman PW. The effect of pressure on the viscosity of forty-three pure liquids. Proceedings of the American Academy of Arts and Sciences. 1926;61:57-99
  38. 38. Gonzalez MH, Lee AL. Viscosity of 2,2-dimethylpropane. Journal of Chemical & Engineering Data. 1968;13:66-69
  39. 39. Kuss E, Pollmann P. Viskositäts-Druckabhängigkeit und Verzweigungsgrad flüssiger Alkane. Zeitschrift für Physikalische Chemie. 1969;68:205-227
  40. 40. Dymond JH, Glen NF, Isdale JD. Transport properties of nonelectrolyte liquid mixtures-VII. Viscosity coefficients for isooctane and for equimolar mixtures of isooctane + n-octane and isooctane + n-dodecane from 25 to 100°C at pressures up to 500 MPa or to the freezing pressure. International Journal of Thermophysics. 1985;6:233-250
  41. 41. Krahn UG, Luft G. Viscosity of several liquid hydrocarbons in the temperature range 298-453 K at pressures up to 200 MPa. Journal of Chemical & Engineering Data. 1994;39:670-672
  42. 42. Pádua AAH, Fareleira JMNA, Calado JCG, Wakeham WA. Density and viscosity measurements of 2,2,4-trimethylpentane (isooctane) from 198 K to 348 K and up to 100 MPa. Journal of Chemical & Engineering Data. 1996;41:1488-1494
  43. 43. Schmidt KAG, Pagnutti D, Curran MD, Singh A, Trusler JPM, Maitland GC, et al. New experimental data and reference models for the viscosity and density of squalane. Journal of Chemical & Engineering Data. 2015;60:137-150
  44. 44. Ciotta F, Maitland G, Smietana M, Trusler JPM, Vesovic V. Viscosity and density of carbon dioxide + 2,6,10,15,19,23-hexamethyltetracosane (squalane). Journal of Chemical & Engineering Data. 2009;54:2436-2443
  45. 45. Canet X, Daugé P, Baylaucq A, Boned C, Zéberg-Mikkelsen C, Quiñones-Cisneros SE, et al. Density and viscosity of the 1-methylnaphthalene + 2,2,4,4,6,8,8-heptamethylnonane system from 293.15 to 353.15 K at pressures up to 100 MPa. International Journal of Thermophysics. 2001;22:1669-1689
  46. 46. Et-Tahir A, Boned C, Lagourette B, Xans P. Determination of the viscosity of various hydrocarbons and mixtures of hydrocarbons versus temperature and pressure. International Journal of Thermophysics. 1995;16:1309-1334
  47. 47. Harris KR, Newitt PJ, Woolf LA. Temperature and density dependence of the viscosity of cyclopentane. Journal of Chemical & Engineering Data. 2004;49:138-142
  48. 48. Kurihara K, Kandil ME, Marsh KN, Goodwin ARH. Measurement of the viscosity of liquid cyclopentane obtained with a vibrating wire viscometer at temperatures between (273 and 353) K and pressures below 45 MPa. Journal of Chemical & Engineering Data. 2007;52:803-807
  49. 49. Hernández-Galván MA, García-Sánchez F, García-Flores BE, Castro-Arellano J. Liquid viscosities of cyclohexane, cyclohexane + tetradecane, and cyclohexane + benzene from (313 to 393) K and pressures up to 60 MPa. Journal of Chemical & Engineering Data. 2009;54:2831-2838
  50. 50. Tanaka Y, Hosokawa H, Kubota H, Makita T. Viscosity and density of binary mixtures of cyclohexane with n-octane, n-dodecane, and n-hexadecane under high pressures. International Journal of Thermophysics. 1991;12:245-264
  51. 51. Jonas J, Hasha D, Huang SG. Self-diffusion and viscosity of methylcyclohexane in the dense liquid region. The Journal of Chemical Physics. 1979;71:3996-4000
  52. 52. Baylaucq A, Boned C, Dauge P, Lagourette B. Measurements of the viscosity and density of three hydrocarbons and the three associated binary mixtures versus pressure and temperature. International Journal of Thermophysics. 1997;18:3-23
  53. 53. Stephan K, Lucas K. Viscosity of Dense Fluids. 1st ed. New York: Plenum Press; 1979
  54. 54. Dymond JH, Robertson J, Isdale JD. Transport properties of nonelectrolyte liquid mixtures - IV. Viscosity coefficients for benzene, perdeuterobenzene, hexafluorobenzene, and an equimolar mixture of benzene + hexafluorobenzene from 25 to 100°C at pressures up to the freezing pressure. International Journal of Thermophysics. 1981;2:223-236
  55. 55. Dymond JH, Awan MA, Glen NF, Isdale JD. Transport properties of nonelectrolyte liquid mixtures. VIII. Viscosity coefficients for toluene and for three mixtures of toluene + hexane from 25 to 100°C at pressures up to 500 MPa. International Journal of Thermophysics. 1991;12:275-287
  56. 56. Ducoulombier D, Zhou H, Boned C, Peyrelasse J, Saint-Guirons H, Xans P. Pressure (1-1000 bars) and temperature (20-100°C) dependence of the viscosity of liquid hydrocarbons. The Journal of Physical Chemistry. 1986;90:1692-1700
  57. 57. Kashiwagi H, Makita T. Viscosity of twelve hydrocarbon liquids in the temperature range 298-348 K at pressures up to 110 MPa. International Journal of Thermophysics. 1982;3:289-305
  58. 58. Roetling JA, Gebhardt JJ, Rouse FG. Effect of pressure on viscosity of naphthalene, phenanthrene and impregnation pitches. Carbon. 1987;25:233-242

Written By

Hseen O. Baled and Isaac K. Gamwo

Submitted: 28 March 2019 Reviewed: 12 April 2019 Published: 07 October 2019