Está en la página 1de 32

PHYSICA

ELSEVIER

Physica A 238 (1997) 81-112

Viscosity of carbon dioxide in the liquid phase


P.S. van der Gulik*
Van der Waals-Zeeman Institute, University of Amsterdam, Valckenierstraat 65-67,
1018 XE Amsterdam, Netherlands

Received 22 July 1996

Abstract

The viscosity coefficient of liquid carbon dioxide has been measured along isotherms from
the triple-point temperature (217 K) up to the critical temperature (304 K) at temperatures of
220,230,240,260 and 280 K by means of a vibrating-wire viscometer. The measurements extended beyond both phase-transition lines into the coexistence region (superheated liquid) and
into the solid range (undercooled liquid). The accuracy of the measurements is estimated to be
1%. The agreement with data of other authors is rather good. For the most part, the results show
a linear pressure dependence in three neighbouring pressure ranges for the various isotherms
with a common intersection with the negative pressure axis pi. The fluidity, the reciprocal of
the viscosity, shows a linear dependence of the molar volume in adjacent density ranges. After
reduction of the molar volume with the volumes of close packing, three sets of linear functions
result with common intersections of the axis VB.
The aim of the present investigation was to produce a new set of viscosity data on request
of the Subcommittee on Transport Properties of the International Union of Pure and Applied
Chemistry in order to provide supplementary data for the improvement of representative equations
for the viscosity of carbon dioxide. The representation of the fluidity as a linear function of the
reduced molar volume in various density ranges provides us with the possibility of presenting
a correlation protocol which produces the viscosity of liquid carbon dioxide within about 1%.
For this purpose, a consistent set of "volumes of close packing" is obtained by means of a
comparison with the results of computer simulations.
The combination of the two types of linear relations gives reason to interpret pi as the internal
pressure and VB as the excluded volume. This interpretation enables us to state that, at constant
temperature, the viscosity coefficient is directly proportional to the thermal pressure, i.e. the
sum of the measured pressure and the internal pressure, and thereby proportional to the number
of collisions of the molecules. The viscosity coefficient is, thus, determined by the number
of collisions per unit time and volume, and by the efficiency of the exchange of momentum
during a collision. From the linearity of the fluidity as a function of the molar volume, we
conclude that the number of collisions is inversely proportional to the free volume V-- VB.

* Tel. +31-20-5256390; fax +31-20-5255788; e-mail: vdgulik@phys.uva.nl.


0378-4371/97/$17.00 Copyright @ 1997 Elsevier Science B.V. All rights reserved
PH S0378-43 71 (96)00466-9

82

P.S. van der Gulik/Physica A 238 (1997) 81 112

Thus our measurements of the viscosity of liquid carbon dioxide lead to simple views on this
phenomenon.
Keywords: Carbon dioxide; High pressure; Low temperature; Liquid; Viscosity coefficient

1. Introduction
In his article "On the Theory of the Friction of Liquids" [1] J.D. van der Waals Jr.
already stated "that the explanation given to account for the friction of gases - viz.
that it is brought about in consequence of this that molecules diffusing from one gas
layer to another, at the same time transport an amount of momentum from one layer
to another - cannot equally apply to the friction of liquids. For the friction of gases
increases at higher temperature. For liquids on the other hand the viscosity becomes
slighter at higher temperature". He continues "for liquids we shall not principally have
to think of transport of momentum by the diffusing molecules, but we shall have to
explain the friction by forces which the molecules exert on each other", and then he
proposes "We might assume that the forces that the molecules exert on each other
at an impact would furnish the explanation of viscosity". However, the theory which
he developed did not differ fundamentally from that for gases, and so the temperature
dependence he found was incorrect. Brillouin [2] remarked that the idea of a mean free
path has to be abandoned for liquids, since they are continuously in a state of collision
with all their neighbours. He therefore developed a theory using elastic waves, but
the resulting viscosity became negative. Andrade [3], in his turn, assumed "that the
viscosity is due to a communication of momentum from layer to layer, but not effected
by a movement of the equilibrium position of molecules from one layer to another,
but by a temporary union at the periphery of molecules in adjacent layers, due to their
large amplitude of vibration". He supposed that the molecules in a liquid vibrate with
the fundamental frequency occurring in the solid state and that this frequency is a
measure for the collision time. To date, his theory is the most successful with respect
to the temperature dependence of the viscosity of liquids.
The classical picture for the viscosity of liquids is therefore that of molecules bouncing around in cages formed by their neighbours, whereby momentum is exchanged
during collisions. The viscosity coefficient is then determined by the number of collisions per unit of time and volume, and by the efficiency of the exchange. On the
other hand, in the spirit of Bernoulli, the pressure is determined by the change of
momentum during collisions of the molecules and is, therefore, also proportional to
the number of collisions. This pressure is, according to Van der Waals Sr. [4], not the
pressure which we measure but the pressure which the molecules experience i.e. the
combination of the measured pressure and the internal pressure due to intermolecular
forces. The latter is generally considered to be a constant background force keeping the
molecules together. In this view we may assume that the magnitude of the momentum,
mv, is solely dependent on temperature, in which case the pressure along an isotherm

P.S. van der Gulik/ Physica A 238 (1997) 81-112

83

is proportional to the number of collisions only, as is also the case for the viscosity
coefficient. Therefore, the viscosity coefficient may be expected to be proportional to
the pressure, which is exactly what we have found in this investigation.
In 1913 Batschinski [5] suggested that at constant pressure the viscosity coefficient
r/ is not only a function of the temperature but especially an inverse linear function of
the specific volume V of the liquid: r / = c / ( V - w), with c and w constant. Later on,
Hildebrand [6] picked up this idea and stated that for constant temperature the viscosity
is an inverse linear function of the free volume V - liB, so r/---- c / ( V - VB) where VB
is the molar volume at fluidity zero ( l / r / = 0). Thus the fluidity, i.e. the reciprocal of
the viscosity, is expected to be a linear function of the molar volume V for isotherms
and, indeed, such linear relations are observed in the present inquiry.
The aim of the present investigation was to produce a new set of measurements of
the viscosity of liquid carbon dioxide on request of the Subcommittee on Transport
Properties of the International Union of Pure and Applied Chemistry in order to provide
supplementary data for the improvement of representative equations for the viscosity
of carbon dioxide which had already been developed [7]. It will be shown that exactly
the representation of the fluidity as linear function of the molar volume in various
density ranges provides us with the possibility of presenting a correlation protocol
which produces the viscosity of liquid carbon dioxide within about 1%.
The measurements of the viscosity coefficient of carbon dioxide in the liquid phase
were performed over the complete liquid range at temperatures of 220, 230, 240, 260 and
280 K, and beyond both phase transition lines into the coexistence region (superheated
liquid) and into the solid range (undercooled liquid) by means of a vibrating-wire
viscometer suited for pressures up to 1 GPa with use of the free damped oscillation
method [8,9]. The details of the experiment are given in Section 2 and the results are
presented in Section 3. In Section 4 we show how a consistent set of "volumes of
close packing" is obtained by means of a comparison with the result of the computer
simulations of Alder et al. [10] and of Michels and Trappeniers [11] using a correlation
developed by Dymond [12]. We do not pretend that these volumes are really the
volumes of close packing, since carbon dioxide molecules are not spherical, whereas
the computer simulation is performed for hard spheres. The reduction of the molar
volumes with this set results surprisingly in coinciding or nearly coinciding values of
VB/Vo for the various fluidity isotherms. Together with the observation that the small
temperature dependence of the slopes is linear and identical for all four ranges and
that the values of the slopes are equidistant, this fact enabled us to construct a simple
correlation function. An explanation for these details is not yet available, however.
By means of this correlation function, it is shown that the agreement of our results
with data of other authors is satisfying, except for the "pioneering" data of Diller and
Ball [13] which are known to be less exact. Finally, in Section 4.4 we show that the
viscosity itself can be expressed as linear functions of the pressure.
In Section 5 an interpretation is given, based on the combination of these two types
of linear relation. This combination results in a simple equation of state, which already
occurs in Brillouin's article [2] and which is reminiscent of the Van der Waals equation

84

P.S. van der Gulik/Physica A 238 (1997) 81 112

of state for low densities. This resemblance encourages us to interpret the intersection
with the pressure axis pi as the internal pressure and VB as the excluded volume. We
would like to stress here that this equation is only valid within the precision of the
present measurements. In the various temperature and density ranges different values
for Pi and VB are found.
In its turn this interpretation of the interception with the pressure axis pi as the
internal pressure leads to the assertion that, at constant temperature, the viscosity coefficient is directly proportional to the thermal pressure, i.e. the sum of the measured
pressure and the internal pressure. This assertion, together with the view that the thermal pressure is proportional to the number of collisions of the molecules, brings us
back to the statement above that the viscosity coefficient is determined by the number of collisions per unit of time and volume, and by the efficiency of the exchange
of momentum. From the linearity of the fluidity as function of the molar volume
we may now conclude that the number of collisions is inversely proportional to the
free volume V - VB, with VB the excluded volume. This statement seems to be quite
plausible.
These assertions do not give a full explanation of the viscosity of liquids, since
the efficiency of the exchange of momentum during a collision is not yet understood.
Our measurements indicate that this efficiency is surely temperature dependent. But
nevertheless it may be stated that our measurements of the viscosity of liquid carbon
dioxide lead to simple views on this matter, in agreement with the early ideas.

2. Method
The viscosity coefficient of carbon dioxide in the liquid phase was measured over
the complete liquid range at temperatures of 220, 230, 240, 260 and 280 K, and beyond
both phase-transition lines into the coexistence region (superheated liquid) and into
the solid range (undercooled liquid). The measurements were performed by means of
a vibrating-wire viscometer suited for pressures up to 1 GPa using the free damped
oscillation method [8,9].
The vibrating-wire viscometer was developed in 1964 by Tough et al. [14]. The
method entails setting a thin tungsten wire into transversal vibration and determining
the damping of this motion by the surrounding sample liquid. The wire had a length of
about 20 mm and a radius of about 25 lam and was set into vibration by means of the
Lorentz force generated by an alternating electrical current and a magnetic field raised
by an electromagnet. After stopping the electrical current, the free damped oscillation of
the wire in the magnetic field causes an induction voltage which is amplified, sampled,
digitalised and stored on a computer disk as a series of 2048 numbers. The damping
of this signal is a measure for the viscosity of the sample fluid.
Pressures lower than 200 MPa were measured by means of a piston gauge, while for
the higher pressures a calibrated manganin cell was used. The pressure vessel, containing the sample fluid and the wire mounted in an insulating support, is part of a cryostat,

P.S. van der Gulik/Physica A 238 (1997) 81-112

85

the temperature of which is regulated to be within a few m K by a proportional-integral


regulation system using a platinum resistance thermometer as sensor. The temperature
was measured by means of a second platinum resistance thermometer using the ITS90.
For the calculation of the density from the temperature and the pressure, the equation
of state obtained from Span and Wagner [15] was applied. The carbon dioxide was
supplied by Air Liquide, Belgium, and was stated by the supplier to be pure within
99.996%.
After the usual cleaning and filling procedure, the vessel was first pressurised at
room temperature, then cooled down to the desired temperature and, subsequently,
depressurised to about 8 MPa. In this way leakage was avoided and extra damping due to insufficient equilibration of the dense liquid was prevented. In the preliminary stage of the investigation, measurements at high pressure immediately after cooling down showed very high damping, afterwards attributed to the inability
of the fluid to transit into a presumably nematic phase during this procedure. The
supposition of the existence of a nematic phase of carbon dioxide at very high densities is also based on the observation that, in a later stage of the investigation, it
was possible to introduce extra damping in a fresh filling at 308 K and 350 MPa
by a severe motion of the wire. This extra damping fades away in a few hours and
is thought to be due to disorder introduced in the nematic liquid by the motion of
the wire.
The usual measuring procedure after equilibration at a certain temperature and pressure was as follows: first a measurement was performed to determine the resonant
frequency of the wire in these circumstances. Then the synthesiser which delivered the
electrical current was tuned to this frequency and, subsequently, five to seven measurements with various sample times were performed. Such a measurement entails the
registration of the sum of one hundred identical signals as obtained by setting the wire
into vibration and sampling the resulting induction voltage. This summation served to
eliminate noise. After Fourier transformation of the resulting set of 2048 stored numbers, the width df and frequency f of the resonance line were determined by matching
the computed signal to a Lorentz function. From the width df, the resonant frequency
f and the density of the fluid, the viscosity coefficient was calculated according to the
method described in Refs. [8,9].
The width df obtained in this way is a measure for the total damping of the signal.
Therefore, extra damping must be avoided, the wire must be clean, the fluid must be
in an equilibrium state and the wire must vibrate at only one frequency. One source of
extra damping cannot, however, be avoided, namely, the mechanical damping of the
wire itself. It was thought that the value of this could be determined in high vacuum,
but the value thus obtained did not agree with the damping measured at atmospheric
pressure and the known viscosity in those circumstances, being too small. Later it
transpired that the value measured after one hour's pumping with a two-stage rotary
oil pump is the correct value; in that case the wire is not yet degassed. For the present
investigation a value was determined from preliminary measurements in vacuum, from
the viscosity of air and of carbon dioxide at atmospheric pressure and by comparing

86

P.S. van der Gulik/Physica A 238 (1997) 81 112

measurements taken at 300 K with those reported earlier [16]. This value must be
subtracted from the computed value of d f / f For the present measurements taken at
high densities the correction is only a few percent.
Subsequently, the results for the different sample times were judged on their quality
and those taken with less suited sample times were consistently left out. The remaining
four or five measurements were averaged to one publishable viscosity value.
The resonant frequency varied from 765 Hz at 280 K, 350 MPa to 823 Hz at 220 K,
0.55 MPa, thus amply satisfying the condition that the frequency must be smaller than
5 MHz. The value of the Stokes parameter m [17] varied from 1.47 at 280K, 350MPa
to 2.79 also at 280 K at the lowest pressure 4.1 MPa, thus satisfying the condition
that 0.14 < m < 6.25 [8,9]. The k-function varied from 1.509 to 1.969, signifying that the hydrodynamic effective mass set in motion by the wire was 1.5-2 times
larger than the mass of the fluid displaced by the wire at rest. The U-function, which
gives the contribution of the friction in the equation of motion, varied from 0.57
to 1.18.

3. Results
The viscosity coefficient of carbon dioxide in the liquid phase was measured at
temperatures of 220,230,240,260 and 280 K as a function of the pressure, which
extended beyond both phase-transition lines into the coexistence region (superheated
liquid) and into the solid range (undercooled liquid). At 280 K the measurements extended only to 350 MPa. The measurements in the metastable states had, as a matter
of course, to be performed rather quickly and, therefore, the accuracy of the measurements is somewhat less than usual: it is estimated to be better than 1%. For
the calculation of the density from the measured temperature and pressure, the equation of state of Span and Wagner [15] was applied, in contrast with the case reported in our interim article [18], where the EOS of Ely [19] was used, since that
of Span and Wagner, which is more accurate and more extensive, was not yet available. Also, previously published results [16] of the viscosity of liquid carbon dioxide
at three temperatures near to the critical temperature, i.e. at 300,303 and 308 K, and
at pressures from 8 to 450 MPa, have been recalculated using the equation of Span
and Wagner [15], since this equation of state is stated to be also valid for pressures higher than 300 MPa and in the metastable states. The temperature range extends from the triple-point temperature (216.58 K) to just over the critical temperature
(304 K).
The experimental results are presented in Tables 1-8. Each table shows the temperature at which the data were measured and the number of data. The first column
gives the pressure at which the data were measured, the second column the mass density calculated from the temperature and pressure with the equation of state of Span
and Wagner [15] and the third column the value of the viscosity coefficient. The data
printed in italics were taken in a metastable state.

87

P.S. van der GuliklPhysica A 238 (1997) 81-112

Table 1
Viscosity of liquid carbon dioxide at the
temperature: 220.005 K, number of data: 17
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(laPa s)

0.559
0.573
0.581
0.581

1166.08
1166.11
1166.12
1166.12

243.4
243.5
243.9
243.6

0.604
1.035
2.528
5.021
7.423
9.727
12.58
13.87
13.87
15.03

1166.18
1167.13
1170.40
1175.67
1180.54
1185.05
1190.43
1192.79
1192.79
1194.89

244.4
243.5
246.8
253.2
257.8
263.5
269.1
272.0
272.9
274.0

17.50
20.05
22.53

1199.24
1203.60
1207.73

279.3
284.0
290.3

Table 2
Viscosity of liquid carbon dioxide at the temperature: 230.005 K, number of data: 26
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(p_Pas)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(lxPa s)

0.872
0.881
0.881

1128.61
1128.63
1128.63

202.7
203.9
203.9

0.900
2.507
5.018
7.477
10.604
14.41
17.84
21.47
24.63
27.05

1128.69
1132.97
1139,34
1145,26
1152,44
1160,68
1167,69
1174,74
1180.58
1184.90

202.6
204.9
210.4
210.9
215.8
223.3
228.7
235.0
240.8
246.4

29.59
32.15
35.40
40.24
45.18
50.22
55.28
60.34
62.64
65.25

1189.29
1193.60
1198.88
1206.42
1213.75
1220.88
1227.71
1234.28
1237.19
1240.41

250.5
254.7
259.2
268.1
275.6
281.7
294.1
302.3
304.0
309.8

67.63
70.31
72.92

1243.30
1246.48
1249.52

313.8
318.0
322.3

T h e e x p e r i m e n t a l results are also s h o w n in Figs. 1 - 3 . In Fig. 1 the v i s c o s i t y coefficient is p l o t t e d as a f u n c t i o n o f t h e density. T h e o p e n s y m b o l s r e p r e s e n t o u r n e w l y


o b t a i n e d results in the stable l i q u i d r a n g e a l o n g the v a r i o u s i s o t h e r m s , a n d the corres p o n d i n g full s y m b o l s r e p r e s e n t t h o s e m e a s u r e d in the m e t a s t a b l e states, b o t h in the

P.S. van der Gulik/Physica A 238 (1997) 81 112

88

Table 3
Viscosity of liquid carbon dioxide at the temperature: 240.004 K, number of data: 32
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(~tPa s)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(~Pa s)

1.255
1,265
1.278

1088.72
1088. 75
1088.80

172.2
172.0
172.0

1.297
2.508
2.514
5.019
7.527
11.05
13.75
17.61
20.78
25.42
28.40
31.45
35.13

1088.86
1092.84
1092.86
1100.71
1108.06
1117.70
1124.61
1133.88
1141.02
1150.87
1156.85
1162.71
1169.46

172.4
175.1
174.6
179.9
182.2
188.3
193.0
199.4
204.1
212.1
216.4
220.0
226.3

40.01
45.18
50.04
55.27
60.33
70.30
80.20
80.20
90.24
100.13
110,22
120,05

1177.98
1186.49
1194.09
1201.87
1209.05
1222.33
1234.53
1234.52
1246.05
1256.72
1266.98
1276.44

232.6
240.2
248.0
255.8
262.6
277.0
290.7
292.2
303.9
317.9
332.0
351.7

124.96
130.44
132.77
135.23

1280.99
1285.93
1288.00
1290.15

359.6
364.9
368.3
374.1

Table 4
Viscosity of liquid carbon dioxide at the temperature: 260.005 K, number of data: 47
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(~Pa s)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(~Pas)

60.33
70.31
80.64
90.62
100.52
110.23
120.67
130.44
140.15
150.46
160.99
170.80
180.57
191.86
201.66
211.45
221.35
230.90
240.60

1158.27
1173,98
1188,80
1202,00
1214,15
1225,32
1236,62
1246,61
1256.04
1265.56
1274.89
1283.10
1291.02
1299.78
1307.10
1314.18
1321.10
1327.56
1333.94

215.5
228.1
244.5
253.6
266.3
278.3
291.5
302.6
314.6
329.1
340.1
353.1
365.0
379.1
389.6
405.0
417.7
434.8
449.9

245.55
250.35
255.10
260.00

1337.13
1340.17
1343.14
1346.16

458.4
463.5
469.2
474.5

2.383
2.402

99&58
998.69

124.7
125.4

2.42l
2.450
2.532
3.413
5.034
6.441
8.119
10.86
14.80
17.38
18.64
21.71
24.74
26.72
28.89
31.45
35.40
35.40
35.40
40.53
45.18
50.22

998.80
998.97
999.42
1004.29
1012.72
1019.51
1027.09
1038.47
1053.07
1061.79
1065.83
1075.17
1083.76
1089.12
1094.73
llOl.08
1110.34
1110.32
1110.31
1121.50
1130.97
1140.59

125.0
124.8
125.7
127.8
130.9
133.2
136.4
141.4
148.1
151.8
154.8
158.9
163.6
167.3
170.6
173.8
179.1
182.0
181.9
190.2
195.7
202.2

89

P.S. van der Gulik/Physica A 238 (1997) 81-112

Table 5
Viscosity of liquid carbon dioxide at the temperature: 280.015 K, number of data: 47
Pressure
(MPa)

Density
(kg/m3)

Viscosity
(I.tPas)

4.094
4.147

882.55
883.25

88.8
88.5

4.192
4.246
5.023
5.766
7.478
11.05
14.44
18.31
21.36
24.63
27.05
29.82
31.68
35.40
40.52
45.17
50.22
60.33
70.31
80.64
80.64
90.62
100.52

883.83
884.70
894.10
902.20
918.36
945.46
965.98
985.62
999.12
1012.16
1021.01
1030.48
1036.51
1047.78
1062.00
1073.81
1085.65
1106.95
1125.50
1142.74
1142.75
1157.90
1171.72

88.8
89.0
91.4
93.5
97.9
105.1
112.0
118.6
123.8
128.4
132.0
136.1
138.8
143.7
150.5
155.8
162.1
174.9
185.8
195.6
197.3
206.6
216.8

Pressure
(MPa)

Density
(kg/m3)

Viscosity
(p.Pa s)

110.22
120.66
130.44
140.16
150.46
160.99
170.79
180.57
191.86
201.66
211.9
221.4
231.1
240.7
250.8
260.3
270.0
279.5
290.7
300.5
319.4
353.2

1184.33
1196.99
1208.12
1218.57
1229.07
1239.25
1248.29
1256.92
1266.44
1274.39
1282.37
1289.52
1296.59
1303.37
1310.28
1316.59
1322.86
1328.84
1335.69
1341.53
1352.38
1370.65

226.6
237.7
247.2
257.0
267.5
278.4
288.5
298.9
310.5
318.6
329.8
340.0
349.6
359.5
370.3
378.5
389.7
400.1
412.9
421.7
443.0
480.9

undercooled liquid state beyond the l i q u i d - s o l i d phase-transition line and in the superheated liquid state beyond the l i q u i d - v a p o r phase-transition line. The recalculated
data obtained at 300, 303 and 308 K are also shown.
In Fig. 2 the same data are plotted as a function of the pressure. Here the various
isotherms can be easily distinguished. Fig. 3 gives an enlarged view of part of the
data shown in Fig. 2 in order to enable comparison with data of Padua et al. [20],
of Diller and Ball [13] and of Ulybin and Makarushkin [21]. The agreement with the
data of Padua et al. [20] taken at 260 K is rather good, with those taken at 280 K
sufficient, but with those taken at 300 K somewhat less. The agreement with the data
of Ulybin and Makarushkin [21], taken at 243.15, 253.15, 273.15 and 293.15 K, is
also satisfactory when the differences in temperature are taken into account. However,
the viscosity values of Diller and Ball [13], taken at the same temperatures as ours,
are generally a few percent too high, as can be clearly observed. A detailed numerical
comparison with the results of other authors is given in Section 4.3.
So far, we have limited ourselves purely to presentation of the results. We will now
analyse the results by comparing them with computer simulations.

90

P.S. van der Gulik/Physica A 238 (1997) 81-112


Table 6
Viscosity of liquid carbon dioxide at the temperature: 300.010 K, number of data: 39
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(p_Pa s)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(~Pa s)

6.792
7.028
7.381
8.023
9.002
10.23
11.53
13.03
14.32
15.36
16.77
17.84
19.37
22.27
25.43
28.13
30.00
40.01
50.27
60.63

687.90
707.76
728.22
753.89
780.96
805.64
825.97
845.09
859.10
869.19
881.54
890.10
901.34
920.14
937.85
951.28
959.84
998.44
1029.63
1055.82

54.2
56.5
59.5
63.2
69.0
73.0
77.7
80.5
83.3
85.4
88.3
90.7
93.1
98.4
103.4
107.5
110.0
123.2
135.3
147.0

80.22
100.10
100.47
115.48
130.81
151.06
175.25
201.66
201.67
226.2
251.9
276.5
301.6
324.0
352.0
376.8
401.7
424.4
453.2

1096.16
1129.28
1129.84
1151.29
1170.93
1194.18
1218.80
1242.71
1242.72
1262.77
1281.98
1298.96
1315.08
1328.60
1344.50
1357.77
1370.43
1381.44
1394.77

166.3
186.1
186.2
199.8
213.6
231.4
253.1
276.6
277.4
297.8
321.2
342.7
364.6
387.0
413.2
439.4
461.5
484.7
514.7

CARBON

DIOXIDE

500

400

0.
300
0

>

200

100

600

800

1000

1200

, I

1400

Density (kglm3 )
Fig. 1. The viscosity coefficient of carbon dioxide as a function of density, open symbols: in the stable state.
filled symbols: in the metastable state. /X 220K, O 230K, V 240K, O 2 6 0 K , [ ] 280K.
300 K, ~C" 303 K and + 308 K.

P.S. van der Gulik/Physica A 238 (1997) 81-112

91

Table 7
Viscosity of liquid carbon dioxide at the temperature: 303.050 K, number of data: 43
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(I.tPa s)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(I.tPa s)

7.232
7.330
7.466
7.623
7.906
8.299
8.801
9.722
10.80
11.98
13.87
15.12
17.84
21.24
24.98
30.00
34.95
40.01
44.94
50.04
59.95
70.31

615.01
640.03
660.55
676.88
698.05
719.03
739.00
765.91
789.21
809.35
834.93
848.91
874.31
899.81
922.90
948.54
969.83
988.78
1005.12
1020.35
1046.30
1069.59

48.3
49.8
51.2
53.7
55.8
58.5
61.3
65.6
69.4
73.4
78.3
81.1
86.8
93.4
98.9
106.0
113.1
119.5
125.4
131.0
142.4
152.4

80.21
90.24
100.13
100.14
115.10
129.88
151.07
175.26
201.66
226.0
250.6
273.1
300.8
324.7
348.7
366.5
378.6
394.5
415.2
435.2
452.0

1089.19
1107.02
1123.01
1123.02
1144.83
1164.15
1188.87
1213.80
1237.99
1258.09
1276.71
1292.49
1310.55
1325.10
1338.87
1348.61
1355.02
1363.20
1373.47
1383.02
1390.78

162.5
172.2
181.7
182.0
193.9
209.7
226.9
249.7
271.5
292.1
313.5
333.4
357.3
380.1
404.3
416.8
431.4
444.2
464.7
480.0
497.8

Table 8
Viscosity of liquid carbon dioxide at the temperature: 308.150 K, number of data: 41
Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(I.tPa s)

Pressure
(MPa)

Density
(kg/m 3 )

Viscosity
(I.tPa s)

8.610
8.845
9.185
9.688
10.32
11.24
12.37
13.87
15.13
15.73
17.85
18.08
21.24
25.10
30.23
35.12
40.01
45.17
50.27
60.34
70.30

626.75
650.08
674.36
700.15
724.06
750.17
774.48
799.51
816.66
824.10
846.61
848.82
875.65
902.04
930.45
952.97
972.41
990.50
1006.47
1033.92
1057.20

47.4
50.3
52.9
55.9
59.2
63.0
67.0
71.5
75.2
76.0
80.8
81.5
87.4
93.7
101.0
108.2
114.3
120.4
126.8
136.2
147.0

80.21
90.24
100.47
101.67
115.48
130.80
151.06
175.25
201.66
226.0
254.1
276.2
300.7
326.3
350.7
351.4
377.2
401.3
424.6
452.5

1077.53
1095.96
1112.99
1114.88
1135.42
1155.88
1179.99
1205.45
1230.10
1250.57
1272.06
1287.64
1303.77
1319.54
1333.66
1334.06
1348.14
1360.63
1372.14
1385.29

156.4
165.6
174.8
175.8
187.8
201.6
218.6
238.9
261.5
282.1
305.3
324.5
346.9
367.2
389.7
390.2
411.2
434.8
456.8
481.8

92

P.S. van der Gulik/Physica A 238 (1997) 81-112

500

X *+

400
0

0O

o0
V

/_,OoO

~*+

o [2

,x * +
+

oo

00

~+

oO

xy v
o
o
~_.~7 0 0 0
o0O
~

--

~
o
oo ~
oOo
oo @ g

200

CARBON DIOXIDE

200

300

400

100

Pressure (MPa)

Fig. 2. The viscosity coefficient of carbon dioxide as a function of pressure, open symbols: in the stable
state, filled symbols: in the metastable state. A 220K, 0 230K, V 240K, O 260K, []
280 K, 300 K, "~r 303 K and + 308 K.

O<>O@@@

~7

300

AA

250

~ A
x~
_@

v~

~ 0
O O

O v
~

00

0 00

^0
O ~

"=

u sv

pv ,,

150

**-.,,*

oO-

~I

*o ~

* =~ 8~
~o

V
~

~vv~

ev ~'~'~

200

*
,

~,o~:j~ o

*o =" ~ D

[g

D +

* +*

" ;

,oo,
50
0

I
20

40

60

I
80

I
100

Pressure (MPa)
Fig. 3. The viscosity coefficient of carbon dioxide as a function of pressure (detail), present data: ~
220K, O 230K, V 240K, O 2 6 0 K , [ ] 280K, 300 K, ~ 303 K and + 308K, open symbols:
in the stable state, filled symbols: in the metastable state.
Diller and Ball [13]: A 2 2 0 K, O 2 3 0 K, [] 240K, [] 260K, n 2 8 0 K and x 300K.
Ulybin and Makarushkin [21]: ~ 2 4 3 K, "~r 253 K and ~ 2 7 3 K.
Padua, Wakeham and Wilhelm [20]: 260 K, 280 K and + 300 K.

P.S. van der Crulik/Physica A 238 (1997) 81 112

93

4. Analysis of the results

4.1. Comparisonwith computersimulations


Due to the lack of a suitable theory, the first step in the analysis of the data is a
comparison with the results of computer simulations. Such a simulation of inter alia
the viscosity of hard spheres as a function of the density was published in 1970 by
Alder et al. [10]. A similar investigation on square-well molecules was reported in
1980 by Michels en Trappeniers [11]; the results of this can easily be extrapolated
to the case of hard spheres. Ignoring Alder's value taken at the melting density, the
following correlation formula was constructed [22], following Dymond [12]:

A(~/V) = {

Vo/V<

1.02

for

1.02 + 15(Vo/V - 0.35) 3

for 0.42 <

Vo/V<

1.02 + 15(Vo/V - 0.35) 3


350(V0/V - 0 . 5 7 5 ) 3

for

Vo/V>

0.575,

0.42,
0.575,

(4.1)

where
~MD = A ( V o / V ) ?/Enskog

(4.2)

Here V0 is the molar volume of close packing of the hard spheres, V the molar volume,
~/MD the viscosity resulting from the computer simulation and r/Enskog the viscosity
resulting from the theory of Enskog for hard spheres. Thus, the result is given relative
to the Enskog value in terms of the relative density Vo/V.In these terms the melting
density is equal to Vo/V= 0.667. Therefore, for hard spheres A(Vo/V) is equal to 1.02
in the first 2 part of the range concerned and increases to nearly the value 2 at the
melting density in the remaining part.
In order to make a comparison the experimental results also have to be taken relative
to r/E,skog and expressed as a function of the relative density. For this purpose, V0 is
taken as an adjustable parameter and the corresponding hard-sphere diameter is used in
the Enskog theory. However, carbon dioxide molecules are not spherical but, according
to computer calculations of the electrical-charge distribution by Kuchta et al. [23],
sphero-cylindrical with an axis ratio of 5. From the Senftleben-Beenakker effect we
know [24,25] that non-sphericity of the molecules increases the viscosity so that, in
order to make a comparison with spheres, a correction factor slightly smaller than 1
must be applied. With some approximations, the theoretical value of this form factor
could be calculated following Kagan en Afanas'ev [26] to be 0.814, where it is assumed
that all collisions are elastic. Therefore, the true value is between 0.814 and 1. The
experimental value of 0.881 obtained from the data measured at 308 K [16] agrees
satisfactorily with this approximation.
In the present analysis the MD results are multiplied with this form factor 0.881:
?/MD = 0 . 8 8 1 A ( ~/0/V)~Enskog

(4.3)

94

P.S. van der Gulik/Physica A 238 (1997) 81 112

13t

CARBON

DIOXIDE

1.2

~ "=' 1.1

IliA

1.0

IIIB
0.9

0.8

0.3

0.4

0.5
Relative Density VoN

0.6

0.7

Fig. 4. The relative viscosity ?]/qEnskog as a function of the relative density Vo/V, open symbols: in the stable
state, filled symbols: in the metastable state: A 220 K, ~ 230 K, V 240 K, O 260 K, []
280 K, 300 K, -~ 303 K and + 308 K.

and both the present and former experimental results are fitted to the resulting curve.
Fig. 4 shows that this procedure leads to a satisfying fit. Most of the values fall in the
last third part of the density range, where the MD result increases to nearly double the
Enskog value. This range III is subdivided into IliA where the value is smaller than
1 and follows Eq. (4.3), into IIIB where the value is larger than 1 and also follows
Eq. (4.3), and into IIIC where the values do not follow Eq. (4.3). The criterion for
this subdivision is, however, given by the behaviour of the fluidity.
The values for the volumes of close packing V0 obtained from these fits are presented
in Table 9 and plotted in Fig. 5. The table shows that they are of the right order of
magnitude, but it must be stressed here again that the term "volume of close packing"
is used for similarity with earlier publications and is not meant to state that this is the
exact value of this volume. The values for our earlier data are slightly lower than those
reported in Ref. [16]. The temperature dependence of V0, shown in Fig. 5, reflects the
fact that, at high temperatures, the molecules move faster, collide with greater impact,
and, therefore, penetrate each other further than at low temperatures.
For comparison with other viscosity data, our own data of V0 appeared to be insufficient for temperatures between 240 and 260 K. Therefore, they are supplemented with
the data of Ulybin and Makarushkin [21] in this temperature range. For this purpose,
these data were analysed in a similar manner. The resulting V0 values are also given
in Table 9 and plotted in Fig. 5. They show clearly that a change in character occurs

P.S. van der Gulik/Physica A 238 (1997) 81-112

95

23

"

>o

21

20
200

I
I
240
280
T e m p e r a t u r e (K)

I
320

Fig. 5. The "volume of close packing" V0 as a function of temperature, * from the present data, [] from
the data of Ulybin and Makarushkin [21].
Table 9
Values of the volumeof close packing V0
Temperature (K)

V0 (m3/Mmol)

Present d~a:
220.005
230.005
240.004
260.005
280.015
300.010
303.050
308.150

22.35
21.88
21.65
21.35
20.95
20.63
20.58
20.50

Ulybin and Makarushkin [21]:


243.15
21.58
253.15
21.43
273.15
21.10
293.15
20.73

around 240 K. As the best fit to the complete set of V0 values we have found:

V0 =

93.4106 - 0.586991T + 0.00119995T 2

for T < 241.5 K ,

2 5 . 7 5 0 8 - 0.0170686T

for T > 2 4 1 . 5 K ,

(4.4)

where V0 is expressed in m3/Mmol and the temperature T in Kelvin. This fit is shown
as a solid line in Fig. 5.
4.2. The fluidity
The second step in the analysis is the reduction of the molar volume with this set of
values of V0, whereupon the fluidity, i.e. the reciprocal of the viscosity, is considered as
a function of the reduced molar volume. The surprising results are shown in Figs. 6-9.

P.S. van der Gulik/Physica A 238 (1997) 81 112

96

C A R B O N DIOXIDE

20

"T

1_.

15

IIIC

i IIIB i

ft.

vE

._=2,
."o_

10

IliA
I

,/Z,''
0
1.00

IIi

15 0

I
2.00

I
2.50

I
3.00

I
3.50

Relative Molar Volume V N o


Fig. 6. The fluidity, the reciprocal of the viscosity coefficient, of carbon dioxide as a function of reduced
molar volume, open symbols: in the stable state, filled symbols: in the metastable state: O 260 K,
[] 280 K, x 300 K, ,k 303 K and + 308 K.

220.005 K

oI
-1 I
1.60

......

I
1.62

I
1.64

I
1.66

xx
I
1.68

I
1.70

230.005 K

=~ 0I] >'0xx
KXX X X X x x X X
.=~'~

-1
1.60

I
1.65

X X

I
1.70

XX
I
1.75

xx

I
1.80

240.004 K

1I
OI
-1 I
1.50

)f(
X
)~1

Xx X

xX

XX XxX
x

XvX xx
x xX "
I
x

1.60
1.70
1.80
Relative Molar Volume V / V o

I
1.90

Fig. 7. Percentage deviations of the fluidity from the linear relations as a function of reduced molar volume
for the present data at 220, 230 and 240 K.

P.S. van der Gulik/Physica A 238 (1997) 81-112


x

01
I
1.70

I
1.60

1.50

260.005 K

:., "X='x

:+

-I

97

I
1.90

I
1.80

i x.~
2.00

1|

I
2.10

280.015 K

0 I :~v.:!~
-1

.....

=xx

...... ,,~

,, .

1.60

1.80

2.00

2.20

2.40

300,010 K
1ix
.t:

x~

01

-1
1.50

X:~--

-- ~ x

: 4 .4.
,+

M ,~

Ix
2.00

XX w

"~

I
2.50
303.050 K

0
-1
1 .50

2.00

11'.~(,dl~
0 i
-1

2.50
308.150 K

,I,I 1 ~

..X~.=4..,."

1,50

-Xx x
I

2.00

X X
x~ ~
x

X X
~

x ~

2.50

Relative Molar V o l u m e V / V o

Fig. 8. Percentage deviations of the fluidity from the linear relations as a function of reduced molar volume,
for the present data at 260, 280, 300, 303 and 308 K. range IIIC, + range IIIB and x range IliA.

In Fig. 6, the fluidity is plotted as a function o f the relative m o l a r v o l u m e for 260,


280, 300, 303 and 308 K. Each isotherm shows three linear ranges, each range with
c o m m o n axis intersection at V~/V0 = 1.405, 1.353 and 1.310, respectively. Thus, in each
range the fluidity behaves according to

1/rl = S t ( T ) . ( V - VB)/Vo ,

(4.5)

as suggested by H i l d e b r a n d [6]. The ranges are n a m e d I l i A , IIIB and IIIC, similar to


earlier papers [27,28,16,18]. The intermediate density range II is reached at 300, 303
and 308K, as can also be seen in Fig. 4. The transition f r o m range II to range IIIA is located at V/Vo = 2.8, Vo/V = 0.357, that f r o m I I I A to IIIB at V/Vo = 1.86, Vo/V = 0.5376
and that f r o m IIIB to IIIC at V/Vo - 1.65, Vo/V = 0.606. The isotherm taken at 260 K
shows transition character with a value o f the intersection with the v o l u m e axis for the
range IIIC o f V~/I~ = 1.318. For each o f the three l o w e s t temperatures, 220, 230 and

98

P.S. van der Gulik/Physica A 238 (1997) 81 112

243.15 K

Oil.

-1
1.70

1.75

1.80

1.85

1,1I

I
1.90

253.15 K
._

1I

.I-

-1

=
,p

I
1.70

1.80

1.90

2.00

273.15 K

0
-1
1.80

Xx

x
x

m
I

I
1.90

x
m ~

2.00

IX
2.10

I
2.20

Relative Molar Volume V/V o


Fig. 9. Percentage deviations of the fluidity from the linear relations as a function of reduced molar volume,
for the data of Ulybin and Makarushkin [21] at 243.15, 253.15 and 273.15 K. + range IIIB and x range IliA.

240K, near the triple-point temperature 216.58 K, one linear function is found, with a
common intersection of the volume-axis at V/Vo= 1.326 and varying slopes.
The deviations from linearity are plotted in Figs.7 and 8 in order to demonstrate how
strictly this linear volume dependence is followed. Fig. 7 shows that the data taken at
220, 230 and 240 K follow the law of linearity within one percent, that is, within the
accuracy of the measurements, with the exception of the higher volume data at 230 K.
This series of six points below 6 MPa was performed separately, four of them were
taken hastily in the superheated liquid. Presumably, the system was not in equilibrium
at that opportunity, the results are 2% too high, as can also be seen in Fig. 3.
The deviations from linearity for the temperatures 260, 280, 300, 303 and 308 K are
shown in Fig. 8. At 260 K, the usual larger deviations at the transition from IliA to
IIIB can be observed, as if the fluid does not know what to do. For the three highest
temperatures there is some uncertainty in the behaviour on the left hand side, i.e. the
points taken at the highest pressures above 300 MPa. It is observed that a better result
can be obtained by splitting up both range IIIB and IIIC, but in view of the number of
points and the accuracy we have not done so. The criterion used is that the deviations
are, on the average, less than one percent. However, this criterion led to the choice of
three ranges instead of two as in the interim article [18].
Finally, Fig. 9 shows the same data for the measurements of Ulybin and Makarushkin
[21 ]. These are taken in a limited density range. It can be seen how the transition from
IliA to IIIB at V/Vo= 1.86 shifts through the data with the temperature. The deviation
from the law of linearity is small, nearly within a half percent; for the data measured
at 293.15 K it is, however, much worse, with deviations up to 7%. These data are
further ignored.

P.S. van der GuliklPhysica A 238 (1997) 81-112

99

Table 10
Values of the axis intersection VB/Vo and of the slope Sf of the
fluidity as a function of the relative molar volume
Temperature (K)
vB/Vo =
220.005
230.005
240.004
vB/Vo 260.005
280.015
3oo.olo
3o3.o5o
3o8.15o

Sf [(mPas) -1]
1.326
11.33
10.97
1o.73
1.31o

1,353

9.832a
9.244
8.775
8.704
8,591

1o.88
10.49
9.972
9.868
9.794

243.15
253.15
273.15
293.15
Valid from V/Vo =
till V/Yo =
range

11.39
11.18

1.65
IIIC

1.65
1.86
IIIB

1.4o5
12.19
11.6o
11.12
11,05
lO,92
12.39
11.85
11.28
1.86
2.80
IliA

a vB/Vo = 1.318.

4.3. Comparison with data o f other authors by means o f a correlation


The linear molar volume dependence o f the fluidity offers the opportunity to construct
a simple correlation fimction for the viscosity results and to compare data o f other
authors with this. The values o f the intersections VB/V0, o f the slopes S f ( T ) and o f the
range limits are collected in Table 10. Just above 240 K, the above mentioned change
in character manifests itself as a splitting into the three ranges IliA, IIIB and IIIC. The
intermediate value at 260K, VB/V0-----1.318, suggests that the low-temperature behaviour
is a continuation o f that in range IIIC. The values o f the slopes S f ( T ) are plotted in
Fig. 10. For the various ranges they show a linear temperature dependence, which is
the same for all ranges, with a slope o f 0.025 ( m P a s K ) - l . Moreover, the values for
the ranges IliA, IIIB and IIIC are equidistant with A S f ( T ) = 1.20(mPas) - t .
The procedure for the calculation o f the viscosity coefficient for given temperature
and density is now carried out according to the following scheme:
a: Vo is calculated from the temperature b y means o f Eq. (4.4),
b: the value o f V/Vo is computed from the density,
c: for T < 241.5 K: 1/r1= (16.76 - 0.025 T) {(V/V0) - 1.326}
according to Eq. (4.5) using the data mentioned above.
For T > 241.5 K: the density range has to be subdivided.

P.S. van der Gulik/Physica A 238 (1997) 81-112

100

13
12
'7,
11
E

CARBONIDIOXIDE

200

240
280
Temperature (K)

320

Fig. 10. The slopes of the fluidity as a function of reduced molar volume plotted as a function of temperature,
for the present data: range IliA, range II1B and range IIIC, and for the data of Ulybin and
Makarushkin [21]: range IliA and range lllB.

for

V/Vo <

St(T)

1.65, range II1C:

= 16.26 - 0.025 T

for T < 2 8 0 K :

VB/Vo =

for T > 280 K :

vB/Vo = 1.31

for

V/Vo <

1.86, range IIIB :

S f ( T ) = ( 1 6 . 2 6 + 1.20) - 0.025 T

VB/Vo =
for

V/Vo <

2.8, range IIIA"

1.326 - 0 . 0 0 0 4 ( T - 2 4 0 )

Sf(T)

1.353

=(16.26

+ 2 . 4 0 ) - 0.025 T

VB/Vo = 1.405

d :

l/. =Sf(T){(V/Vo)

(VB/Vo)}

(4.6)

Sf(T) is expressed in (mPa s) -1.


For V/Vo > 2.8, range II, it is not possible to construct a correlation based on the
present data.
Due to minor differences, mainly in the slopes Sf(T), the deviations of the present
data from the correlation given above are, of course, larger than in the case of Figs. 7
and 8, but in general the behaviour is the same. Only the data taken at 260 K in
the density range IIIB deviate systematically more than 1% from the correlation, due
to the choice of a different slope. This choice is based on the results of the analysis of the data of Ulybin and Makarushkin [21], as can be observed in Fig. 10,
whose analysis was included in order to remove the uncertainty in the behaviour of
the fluidity in between 240 and 260 K. Of course, the correlation is consistent with
their data.
The correlation now offers the opportunity of comparing the results with those of
Golubev and Shepeleva [29] which were not taken along isotherms, but also cover
the low temperature region. The densities are calculated from the given pressure and
temperature by means of the equation of state of Span and Wagner [15] and, consequently, the viscosity is calculated by means of the correlation from these densities
and the corresponding temperatures. The deviations of the experimental values from

P.S. van der Gulik/Physica A 238 (1997) 81-112

Golubev

.'<
x

xx

-2

-3
I
900

-4 ,
800

I
1000

I
1100

I
1200

Padua
4.

o~
.=_

y.

x~x.x

-1

101

t $,+$ +

..

2
1
0
-1
I

-2
8O0

1000

1100

1200

Michels
o

-2

I~

-3 ~-

-4
800

I
900

I
1000
Density in kg/rn 3

&

I
1100

I
1200

Fig. 11. The percentage deviations of the viscosity coefficient of carbon dioxide from the correlation as a
function of density, for data of Golubev and Shepeleva [29], x, at temperatures from 241.75 to 293.85 K,
for the data of Padua et al. [20], at 260 K, I~ at 280 K and + at 300 K and for data of Michels et al.
[30], ~ at 298.15 K, V at 303.05 K, at 304.25 K, [~ at 305.15 K, A at 313.15 K, at 323.15 K and rl at
348.15 K.

the calculated values are shown in Fig. 11. In general, the data show a non-systematic
spread of 2% around the correlation values. Especially the results obtained at low
temperature justify our choice of correlation in this temperature range.
The data of Diller and Ball [13] deviate at all temperatures from the correlation given
above, on average + 3 % at 220 K, + 8 % at 230 and 233 K and + 4 % at 240, 260 and
280 K. Other data taken at these low temperatures were not available. Deviations from
data taken by Padua et al. [20] are also shown in Fig. 11. The agreement is perfect
for those taken at 260 K, less perfect for those taken at 280 K and still worse for
those taken at 300 K, where a deviation of on average + 3 % is observed for the lower

102

P.S. van der Gulik/Physica A 238 (1997) 81-112

densities. Like our data both those of Diller and Ball [13] and those of Padua et al. [20]
are measured by determining the damping of an oscillator. With this method, the ideal
damping due to the viscosity alone is the minimum one, any disturbance producing an
increase of the damping and therewith an apparent value of the viscosity coefficient
which is too high. This is also true for our low density data at 230 K.
Finally, the deviations of the "old" data of Michels, Botzen and Schuurman [30]
from the present correlation are also plotted in Fig. 11. They show negative deviations
in the order of one percent, due to the manner in which the viscosity coefficient was
calculated from the experimental observations in the pre-computer era, as explained in
a paper by Van den Berg et al. [31].
For the lower densities in range II at temperatures around room temperature, the
agreement of the present results with the data of Michels et al. [30] and those of
Iwasaki and Takahashi [32] is also satisfying.
4.4. The pressure dependence o f the viscosity coefficient

The success of the analysis of the fluidity of carbon dioxide in terms of linear functions of the molar volume gives rise to a closer examination of the pressure dependence
of the viscosity coefficient, which was earlier found also to be linear in argon [33] and
methane [28]. Fig. 12 shows that it is indeed possible to discern ranges where the
viscosity coefficient is a linear function of the pressure:
r1 = z ( T ) { p +

pi}.

(4.7)

r(T) is the slope of the lines, expressed in picoseconds, and Pi the negative value of
the intersection with the pressure axis, expressed in MPa.
For temperatures near the triple-point temperature (216.58 K) up to 240 K, a linear
pressure dependence of the complete viscosity isotherm is observed, as in the case
of the fluidity, with a coinciding intersection of the pressure axis, P i - - l l 6 M P a . The
deviations from linearity are plotted in Fig. 13. In general, the spread is one percent,
being the estimated accuracy, but again the low density data at 230 K are seen to be
too high. The data for 240 K already show a tendency to split up into three ranges, as
observed for the higher temperatures.
For the temperatures higher than 240 K, a linear relation is observed for the corresponding viscosity range, marked M in Fig. 12, with an intersection with the pressure
axis which gradually shifts towards lower values with the temperature. This range extends from r/= 185 to 400 laPa s. In this case, the ranges are not limited by pressure
values but by viscosity values, in contrast with the case of fluidity, where the ranges
are molar volume ranges and not fluidity ranges. Therefore, the ranges of linear volume
dependence of the fluidity can never coincide exactly with the ranges of linear pressure
dependence of the viscosity.
For viscosity values higher than 400 laPa s, which occur only at temperatures above
240 K, a linear pressure dependence can again be observed, now with a coinciding
axis intersection of about 75 MPa. They are indicated in Fig. 12 by the dashed lines,

P.S. van der Gulik/Physica A 238 (1997) 81-112

500

i//

"~

400

.i

300

103

zf
. . . .

....

;~///

8
200
"L'

. . . . . . . . . . . . . .

100
CARBON DIOXIDE

IlVt
- 100

I
0

I
100

I
200

I
300

Pressure(MPa)

I
400

I
500

Fig. 12. The viscosity coefficient of carbon dioxide as a function of pressure, open symbols: in the stable
state, filled symbols: in the metastable state. /k 220 K, O 230 K, V 240 K, O 260 K, []
280 K, 300 K, ,~ 303 K and + 308 K.

220.005 K
0
-1
0

10

15

20

2 i~.,. +
"
'-~"

I"
0 I
-1 I
0

230.005 K
. "=.
-=-,1, .L- "="

a. "1"+'1"

.=. -I-

_;_

I
25

I
35

4,

"i"i"l'-II
75

240.004 K

4,

ilk
50

1 I @ 4.4"4- v + @ + @ +
-1

25

I
-I70
Pressure (MPa)

+@
dk I
105~"

+
I
140

Fig. 13. Percentage deviations of the viscosity coefficient from the linear relations as a function of pressure
for the present data at 220, 230 and 240 K.

P.S. van der Gulik/Physica A 238 (1997) 81 112

104

the range is indicated with H. The existence of this range is in agreement with the
analysis of the methane data [27,28], which also showed such a linear high pressure
range.
For viscosities lower than 185 pPa s, which also occur only at temperatures above
240K, such a linear behaviour can be discerned too. The range is limited to the viscosity
values between 135 and 185 pPas and is indicated with L in Fig. 12. The intersection
with the pressure axis occurs at about 84 MPa, the slopes r are temperature dependent,
as in any range.
The range of viscosities smaller than 135 pPas is indicated with S. In this range the
viscosity isotherms taken above the critical temperature are S-shaped. Of course, an
S-shaped curve can be approximated by linear line-pieces over small ranges, but these
are not considered to be of importance. The choice of the number of ranges and their
limits is always rather arbitrary. As in the case of the fluidity, deviation plots such as
Figs. 13 and 14 have been taken as the criterion. In Fig. 14 it can be observed that, in

260.005 K

0
-1
0

1
0 [
J

50

100

150

200

250
280.015 K

~.x..~,,I-...~,_,.
X ~" It . . . .
~-,-.-

-1

~ . , ~ ~.XX
,e,-rr X X

100

200

300

400
300.010 K

1 I

<1

-1

x
X

t,

_.,,....,,. "1"

~-=-I

-,- -

1O0

- -iI

a.

- ~

200

X X
I

300

~..

X
I

X
I

400

500

303.050 K
1 I
0 I

-1

+
" - ''*.,.e
-~

I +

+
~.

+.x
%

100

-I.

200

300

-I

X ,,~

400

1[
0

x
""

500

308.150 K
xxx_~

wI

1O0

200

",-

Pressure

300
(MPa)

.6

vX

~ ....

400

500

Fig. 14. Percentage deviations of the viscosity coefficient from the linear relations as a function of pressure
for the present data at 260, 280, 300, 303 and 308 K. x range L, + range M and range H.

P.S. van der Gulik/Physica A 238 (1997) 81-112

105

Table 11
Values of the axis intersection Pi and of the slope z of the viscosity as a function of pressure
Temperature Pi
z
(K)
(MPa) (ps)
220.005
230.005
240.004

116
116
116

260.005
280.015
300.010
303.050
308.150

2.091
1.711
1.484

135 a < ~ < 185


range L

185 < g < 400


range M

~ > 400
range H

Pi
z
(MPa) (ps)

Pi
r
(MPa) (ps)

Pi
z
(MPa) (ps)

114
112
110
106
104

75
75
75
75
75

84
84
84
84
84

1.510
1.207
1.012
0.988
0.948

1.239
1.020
0.887
0.882
0.855

1.419
1.126
0.970
0.946
0.913

a~/ in pPas.

120

t13

100

I1.

Q.

80

.---.--g-.
CARBON DIOXIDE

60
200

I
I
240
280
T e m p e r a t u r e (K)

I
320

Fig. 15. The intersections of the linear viscosity-pressure relations, interpreted as the internal pressure, plotted
as a function of temperature.

general, the deviations are within one percent, which is estimated to be the accuracy
of the measurements, with the exception of a fluke at 260 K.
The values of the intersections Pi, of the slopes z(T) and of the range limits are
collected in Table 11. In Fig. 15 the values of the intersections are plotted as a function
of the temperature with the division into the three ranges L,M and H above 240K. The
low-temperature intersection is clearly a continuation of that of range M. The values
of the slopes z(T) are plotted in Fig. 16, the lines are meant as a guide for the eye.
Thus far the data have only been analysed without any interpretation. A discussion
of the data will be given in the following section.

P.S. van der Gulik/Physica A 238 (1997) 81-112

106

IARBONDIOXIDE

2.00
1.50
1.00
I

200

240

280

Temperature (K)

320

Fig. 16. The slopes of the linear viscosity~ressure relations, interpreted as the relaxation time, plotted as a
function of temperature.

5. Discussion

5.1. Interpretation
In an attempt to interpret the results obtained thus far Eqs. (4.5) and (4.7) are
multiplied, whence

{p + pi}{V--VB}/Vo = 1/z(T)Sf(T).

(5.1)

This equation of state is derived from the analysis of the linear molar volume
dependence of the fluidity and the linear pressure dependence of the viscosity and
can, therefore, not be more accurate than this analysis. The problem whether experimental pVT data satisfy this equation of state in some density ranges will not be
considered here. We are looking for an interpretation of Pi, VB/V0, z(T) and Sf(T).
Eq. (5.1) suggests that Pi can be interpreted as the internal pressure. The total
pressure applied to the molecules, the thermal pressure, is then equal to p + pi. The
positive value of pi points to the existence of attractive forces between the carbon
dioxide molecules in the liquid state. In the case of argon [22,33] and methane [27,28]
zero and negative values, pointing to the absence of intermolecular forces and repulsive
forces, respectively, have also been found at high densities, but at high temperatures
with respect to the critical temperature. Measurements of the viscosity at high densities
and at low temperature on these materials will show if the absence of repulsive forces
for the present data is due to the temperature or is a property of carbon dioxide.
The internal pressure can be calculated from the equation of state of Span and
Wagner [15] for carbon dioxide. The values obtained in this way are higher than those
of pi and are not constant. At a density of 1185 kg/m 3 they vary gradually from
377 MPa at 220 K to 325 MPa at 300 K, at 220 K they vary from 364 MPa at the
lowest fluid density to 391 MPa at the highest and at 300 K they vary from 84 MPa
at the lowest fluid density to 297 MPa at the highest with a maximum of 346 MPa at
a density of 1260 kg/m 3. This maximum does not occur for temperatures lower than
240 K but is present at higher temperatures at nearly the same density and at pressures

P.S. van der GuliklPhysica A 238 (1997) 81-112

107

of about 200 MPa. So the calculated values are at least of the same magnitude as Pi
and, in general, show the same behaviour.
With this interpretation of Pi, the viscosity of the liquid is directly proportional to
the thermal pressure. If the attractive forces between the molecules can be considered
as being due to a background potential which holds the liquid together, the thermal
pressure is due to the kinetic momentum exchange during collisions and is proportional
to the number of collisions. The linear relation between viscosity and pressure at constant temperature then implies that the viscosity is directly proportional to the number
of collisions, i.e. to the number of opportunities for exchanging momentum. This seems
to be fairly logic for momentum transport. However, it contrasts with the behaviour
of the viscosity at low densities where the mean-free-path approximation is valid. At
these low densities the molecules are no longer locked in by their neighbours and can
pass each other, so that the concept of a mean-free-path length becomes meaningful.
The momentum is then transported by travelling molecules and, on collision, simply
handed over. In this way the viscosity becomes independent of density, pressure and
thus number of collisions in first approximation.
Traditionally, {V--VB} is interpreted as the free volume, i.e. the volume available for
the molecules to move in, and thus VB as the excluded volume. For binary collisions
only this excluded volume is the covolume of Van der Waals b. That the magnitude of
VB is about half that of b points, therefore, to many multiple collisions. On the other
hand, VB is 1.3--1.4 times the volume of close packing due to the disorder in a fluid.
At a molar volume equal to VB, the fluidity becomes zero and the viscosity infinite,
therefore VB must be smaller than the volume at the glass transition. Va is also smaller
than the molar fluid volume at melting. Therefore, in an experiment the volume VR can
never be reached by a liquid.
On decreasing density, we found transitions from small excluded volumes to larger
excluded volumes. This has also been found for argon [22,33], where two values could
be distinguished, and for methane [27,28], where also three different values were found.
In the last case the high-density transition can be seen as the transition from interlocked
rotation of the molecules to free rotation of the molecules [27] and such a transition
indeed produces a larger excluded volume. For the present case, it is possible to see the
high-density transition as the transition from a nematic phase, as mentioned in Section 2,
to the disordered fluid. Also, such a transition produces a larger excluded volume. The
deviation of the experimental data in range IIIC from the computer simulation, as
observed in Fig. 4, also points to such a jump in the value of V0, just as in the case
of methane. If carbon dioxide is really in a nematic phase in the density range IIIC,
then it follows that it would be nematic at all densities at temperatures below 240 K,
since the behaviour at these temperatures is a continuation of that in range IIIC at
higher temperatures. In that case, the deviating results at 230 K can be understood as
being due to the inability of the system at that moment in that particular experimental
situation to transit into the nematic phase.
According to Eq. (4.5), the fluidity is proportional to the free volume {V-VB}. In
the simple view developed above, the thermal pressure is due to the kinetic momentum

108

P.S. van der Gulik/Physica A 238 (1997) 81-112

exchange during collisions and is proportional to the number of collisions. The velocity
of the molecules is then only dependent on the temperature. When the free volume is
enlarged with some factor u at constant temperature, the distance for a molecule to
move freely, for instance in the x-direction, is enlarged with the cubic root of u and,
therefore, the number of collisions for molecules moving in the x-direction is decreased
with the cubic root of u. The same is valid for movement in the y and z direction,
so that the total number of collisions is diminished with the factor u. The number
of collisions is, therefore, inversely proportional to the free volume and thus to the
fluidity. Since the fluidity is the reciprocal of the viscosity, the viscosity is, therefore,
proportional to the number of collisions. This is in full agreement with the view given
above.
Besides the small temperature dependence of V0, the effects of the temperature in
Eq. (5.1) are concentrated in the fight hand side term 1 / z ( T ) . S f ( T ) . V o / z ( T ) . S f ( T ) is
thus the kinetic energy term. Its value ranges from about 0.5 RT per mol at 220 K to
RT per mol at 308 K and at 260 K from 0.5 RT per mol at the low-density side, the
combination of the ranges IIIA and L, to about 0.7 RT per mol at the high-density
side, the combination of the ranges IIIC and H. At 308K these values are 0.77RT per
mol and RT per mol, respectively.
The interpretation of Sf(T) is still difficult. In the simple view developed above it
determines the linear increase of the fluidity with the increase of the molar volume due
to the decrease of the number of collisions at constant temperature. The temperature
dependence of S f ( T ) is, however, not a simple inverse square root dependence due to a
square root temperature dependence of the velocity of the molecules, but a linear one,
as shown in Fig. 10. Apparently, the "efficiency" in exchange of momentum during a
collision and its temperature dependence plays a role.
As will be shown in a forthcoming article, z ( T ) is the time between collisions and the
relaxation time of velocity fluctuations. Its temperature dependence, plotted in Fig. 16,
is not yet understood. It also does not show the square root temperature dependence of
the velocity; on the contrary, when this temperature dependence is taken into account,
the efficiency of the momentum exchange at a collision at 220 K is twice as large as
at 300 K. If the magnitude and temperature dependence of z(T) were understood, it
would be possible to produce a valid theory for the viscosity of dense fluids.
5.2. Conclusions

This paper presents new measurements of the viscosity coefficient of carbon dioxide
in the liquid phase from the triple-point temperature (216.58 K) up to the critical
temperature (304 K) over the complete density range passing both phase-transition
lines. From the analysis of these measurements, a simple view on the viscosity arises
in accordance with classical concepts. In the various ranges viscosity is proportional
to thermal pressure and, therefore, to the number of collisions between molecules.
This number is inversely proportional to the available free volume V--VB, leading to a
linear relation between the fluidity and the molar volume. The value of the viscosity

P.S. van der GuliklPhysica A 238 (1997) 81-112

109

CARBON DIOXIDE
500

400
/)

O3

O.
::L

300

.m
O

._

>

2OO
v

L
lOO-

0
0.3

I
0.4

IIIC

IIIB

IliA
II

I
0.5

I
0.6

0.7

Relative Density Vo/V


Fig. 17. A survey of the various ranges in a plot of the viscosity coefficient of carbon dioxide as a function
of relative density, open symbols: in the stable state, filled symbols: in the metastable state. A 220 K,
(> 230K, V
240K, O 2 6 0 K , [ ] 280K, 3 0 0 K , - k 3 0 3 K a n d + 308K.

coefficient of simple liquids at a certain temperature is then determined by the number


of collisions per unit time and volume and the temperature-dependent efficiency in
momentum exchange during a collision.
A survey over the various ranges is presented in Fig. 17. The viscosity coefficient
is plotted as a function of the relative density Vo/V. It can clearly be seen that more
data are covered by the linear fluidity ranges IIIA, IIIB and IIIC than by the linear
viscosity ranges L,M and H. Therefore, we preferred to use the linearity of fluidity
for the correlation given in Section 4. The data measured at liquid densities could be
presented within 1% by this correlation. It is observed that the transition from range L
to range M more or less coincides with the transition from range IIIA to range IIIB.
This is not true for the other transitions.
Finally, in Figs. 18 and 19 it can be seen which ranges are covered in the p-T
diagram. The triple point and the critical point are indicated with black circular dots
and the melting line and the saturated-vapour-pressure line by solid lines. It is clear
that some measurements have been performed in the metastable region over the phase
lines. Dashed lines indicate the transitions to the various ranges, IIIC, IIIB, IIIA and
II in Fig. 18 and H, M, L and S in Fig. 19. The transition to range II in Fig. 18 is
only given near to the critical point. By comparison of the two figures it can again be
observed that the transitions from range IIIA to range IIIB and from range L to range
M coincide.

P.S. van der Gulik/Physica A 238 (1997) 81 112

l l0

500

/
CARBON DIOXIDE

zx~ zx

400
/*,
13.

300
Solid

"-'1
/1
t/)

,=

/~
/ ~

200

13.

Z~
A

"A

Z~

",,. ~

a~

~"
."

z~ A
A A

oo

oo

ooo

.'8

IIIB

oo

100-

0
200

220

240

260

280

300

320

Temperature (K)
Fig. 18. A survey of the ranges covered by the present data in a p - T diagram of carbon dioxide.
The triple-point and the critical point are indicated by the black dots, the melting line and the saturated-vapour-pressure line by the solid curves, and the three ranges where the fluidity is a linear function of
the molar volume by the open symbols.

500

CARBON DIOXIDE

zxzx A

400

ZkZX

,,

AZ~

Z~.s

^/

#
,,
.~.. "

oo

-~

_/

~
~

~
~

240

260

280

300

/
0.

300
Solid

/~.-

I"1

z~ ..

AO O
o

oo o

200 "
a..

/.-

ooo
o

100

0
200

220

320

Temperature (K)
Fig. 19. A survey of the ranges covered by the present data in a p - T diagram of carbon dioxide.
The triple-point and the critical point are indicated by the black dots, the melting line and the saturated-vapour-pressure line by the solid curves, and the three ranges where the viscosity coefficient is a
linear function of the pressure by the open symbols.

P.S. van der Gulik/Physica A 238 (1997) 81-112

111

This investigation was undertaken at the request of the Subcommittee of the IUPAC
Commission 1.2 on Transport Properties of Fluids in order to provide supplementary
data for the improvement of representative equations for the viscosity of carbon dioxide
which had already been developed [7]. For liquid, in effect this equation consists of
one linear relation between fluidity and molar volume. The deviations of our data from
this equation amount to 7% within its validity range and up to 20% outside this range
at high pressures. The application of the correlation given above is therefore a large
improvement.

Acknowledgements
The author wishes to thank Drs M. E1 Kharraz for his assistance in performing the
measurements and Drs B. E1-Boubkari for his assistance in the analysis of the various
special phenomena which were encountered during this investigation.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]

[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]

J.D. van der Waals Jr., Proc. Kon. Akad. Wetensch. Amsterdam 21 (1919) 743, 1283.
L. Brillouin, J. de Phys. (Paris) 3 (1922) 326.
E.N. da C. Andrade, Phil. Mag. 17 (1934) 497, 698.
J.D. van der Waals Sr, Doctoral Dissertation, Leiden (1873).
A.J. Batschinski, Z. Phys. Chem. 84 (1913) 643.
J.H. Hildebrand, Viscosity and Diffusivity (Wiley, New York, 1976).
V. Vesovic, W.A. Wakeham, G.A. Olchowy, J.V. Sengers, J.T.R. Watson and J. Millat, J. Phys. Chem.
Ref. Data 19 (1990) 763.
R. Mostert, P.S. van der Gulik and H.R. van den Berg, Physica A 156 (1989) 909.
P.S. van der Gulik, in: Experimental Thermodynamics, Vol III: Measurement of the Transport Properties
of Fluids. p. 79; W.A. Wakeham, A. Nagashima and J.V. Sengers (Eds.), IUPAC (Blackwell Scientific
Publications, Oxford, 1991).
B.J. Alder, D.M. Gass and T.E. Wainwright, J. Chem. Phys. 53 (1970) 3813.
J.P.J. Michels and N.J. Trappeniers, Physica A 104 (1980) 243.
J.H. Dymond, Physica A 79 (1975) 65.
D.E. Diller and M.J. Ball, Int. J. Thermophys. 6 (1985) 619.
J.T. Tough, W.D. McCormick and J.G. Dash, Phys. Rev. 132 (1963) 2373. Rev. Sci. Instr. 35 (1964)
1345.
R. Span and W. Wagner, J. Phys. Chem. Ref. Data 25 (1996) 1509.
P.S. van der Gulik, R. Mostert and H.R. van den Berg, High Temperatures-High Pressures 23 (1991) 87.
Sir G.G. Stokes, Mathematical and Physical Papers, Vol. 3 (Cambridge University Press, Cambridge,
1901 ).
P.S. van der Gulik and M. El Kharraz, Int. J. Thermophys. 16 (1995) 145.
J.F. Ely, J.W. Magee and W.M. Haynes, Thermophysical properties for special high CO2 content
mixtures, GPA Research Report RR110 of Project 839 Part I (1987).
A. Padua, W.A. Wakeham and J. Wilhelm, Int. J. Thermophys. 15 (1994) 767.
S.A. Ulybin and V.I. Makarushkin, Thermal Eng. 23 (1976) 65.
P.S. van der Gulik and N.J. Trappeniers, Physica A 135 (1986) 1.
B. Kuchta, R.D. Etters and R. LeSar, J. Chem. Phys. 97 (1992) 5662.
H. Senftleben and H. Gladisch, Ann. Physik 30 (1937) 713.
F. Zernike and C. van Lier, Physica 6 (1939) 961.
Yu. Kagan and A.M. Afanas'ev, Sov. Phys. JETP. 14 (1962) 1096.

112

P.S. van der GuliklPhysica A 238 (1997) 81-112

[27] P.S. van der Gulik, R. Mostert and H.R. van den Berg, Physica A 151 (1988) 153.
[28] P.S. van der Gulik, R. Mostert and H.R. van den Berg, Fluid Phase Equilibria 79 (1992) 301.
[29] I.F. Golubev and R.I. Shepeleva, Kimiya i teknologiya organicheska sinteza, ONT1, GIAP, Part 8
(1971) 44.
[30] A. Michels, A. Botzen and W. Schuurman, Physica 20 (1954) 1141.
[31] H.R. van den Berg, C.A. ten Seldam and P.S. van der Gulik, Physica A 167 (1990) 457.
[32] H. lwasaki and M. Takahashi, J. Chem. Phys. 74 (1981) 1930.
[33] N.J. Trappeniers, P.S. van der Gulik and H. van den Hooff, Chem. Phys. Lett. 70 (1980) 438.

También podría gustarte