Properties of organic-water mixtures. 13. Self-diffusion coefficients of

Apr 1, 1977 - Harold O. Phillips, Arthur J. Shor, Arthur E. Marcinkowsky, Kurt A. Kraus. J. Phys. Chem. , 1977, 81 (7), pp 682–683. DOI: 10.1021/j10...
0 downloads 0 Views 253KB Size
682

Kraus et al.

New York, N.Y., 1967, Chapter 23, p 794. (15) We thank Mr. John Csurny for making these viscosity and density measurements. (16) R. H. Stokes and I.A. Weeks, Aust. J. Chem., 17, 304 (1964). (17) See, for example, (a) R. H. Stokes and R. Mills, “Viscosity of Electrolytes and Related Properties”, Vol. 3, Topic 18 of ‘The International Encyclopedia of Physical Chemistry and Chemical

(18) (19) (20) (21)

physics”, Pergamon Press, New York, N.Y., 1965. (b)R. E. Meredkh and C. W. Tobias, “Advances in Electrochemistryand Electrochemical Engineering”, Vol. 2, Interscience, New York, N.Y., 1962, p 15. R. Simha, J. Phys. Chem., 44, 25 (1940). H. Fricke, Phy. Rev., 24, 575 (1924). J. H. Wang, J Am. Chem. Soa, 76, 4755 (1954). D. A. G. Bruggeman, Ann. Phys., 24, 636 (1935).

Properties of Organic-Water Mixtures. 13. Self-Diffusion Coefficients of Na’ in Glycerol Triacetate-Water Mixtures’i2 Harold 0. Phlllip~,~ Arthur J. Shor, Arthur E. Marcinkow~ky,~ and Kurt A. Kraus” Chemistry Division, Oak Ridge National Laboratory, Oak Ridge, pnnessee (Received November 9. 1976) Publication costs assisted by the U.S. Energy Research and Development Administration

Self-diffusion coefficients,D, of Na” in water solutions of glycerol triacetate (GTA) were obtained at 25 O C as a function of water fraction, f,,and sodium perchlorate concentration. The latter was used to produce miscibility of water and GTA over the composition range studied. At constant NaC104concentration,D decreases rapidly with water content, qualitatively as expected from the viscosity of the solutions. A plot of log D vs. log f,is essentially linear, with a slope of 1.5. The diffusion coefficient-viscosity products are not quite constant but have a shallow maximum near f,= 0.5. A t constant organic volume fraction of 0.97, an increase in NaC104 concentration causes a large increase in the viscosity of the solutions and a somewhat smaller increase in diffusion coefficients.

Our earlier thermodynamic studies of glycerol triacetate (GTAI-water mixtures5 were carried out because we believe that this system, as well as similar ones, are reasonable models for the thermodynamic (equilibrium) properties of cellulose acetate membranes. While one would not expect a water-organic mixture containing a low molecular weight organic compound to be a good kinetic model for a membrane, it seemed nevertheless of interest to examine the diffusional properties of the GTA-water system and, perhaps, even to establish how large the difference between it and a membrane system is. GTA and water are not miscible in all proportions. This immiscibility makes it impossible to examine the diffusion properties over a broad composition range. Since GTA and water can be made miscible in all proportions through addition of modest quantities of certain salts: we elected to make most of our measurements in GTA-water mixtures containing 1.4 M of NaC104 (moles per liter of the GTA-H20 solvent). Since sodium perchlorate has a significant effect on the viscosity of these systems, we have also carried out some measurements as a function of NaC104concentration in GTA solutions of very low water content.

Experimental Section Method. The radiometric porous frit method which was used for measuring the self-diffusioncoefficients has been previously described.’ Briefly, the method consists of saturating a porous frit with the solution of interest containing a radioactive isotope of the ion of interest and measuring the rate of removal of the radioisotope as a solution of identical composition (not containing the tracer) is pumped past it. The diffusion coefficients can then be calculated from the diffusional decay half-times obtained in a flow regime where the measured half-times of removal are essentially independent of flow rates past the frit. Corrections for radioactive decay are made when necessary. The Journal of Physlcal Chemistry, Vol. 81, No. 7, 1977

TABLE I: Self-Diffusion Coefficients of Na’ in Glycerol Triacetate-Water-NaC04 Mixtures at 25 C (1.4M NaC0,) Volume lo”, fraction H’O (f”) shob cm’ls DslD,s,

0.934a 0.844 0.721 0.469 0.291 0.148 0.045

1.00 1.61 2.38 4.94 10.3 25.3 111.6 aNoGTA. b q = 0.952CPfor

1.26 1.00 1.03 1.32 0.756 1.43 0.415 1.63 0.196 1.60 0.074 1.49 0.016 1.42 1.4 M NaClO, solution.

Porous gold frits with about 45% void volume were used and calibrated as described earlier.’,’ In separate experiments, it was shown that the gold frit did not adsorb sodium ions significantly from the GTA-water mixtures. The temperature of the measurements was 25 i 0.1 “C, except for the more viscous solutions where variations were f0.2 “C. Materials. The tracer 24Na(TI,, = 15.05 h) was obtained from the Isotopes Division of ORNL and used without further purification since examination of the y-energy spectra showed no detectable impurities. Use of the tracer was discontinued after several half-lives as a precaution against significant relative buildup of possible long-lived contaminants. The solutions were prepared from reagent grade NaC104 and Eastman Kodak glycerol triacetate. Each solution was checked for water content by Karl Fischer titration. Viscosities were measured by standard techniques (Cannon-Ubbelohde viscometer). Results and Discussion The self-diffusion coefficients of Na” were measured in GTA-H20 solutions as a function of solution composition.

Properties of Organic-Water Mixtures 0

”I

Ln

L

-

0.10

, ,

< 0.06 \ - 0.02 O*O4I 0

l

I

l

l

1

1

,

,

I

1

(A)

1

U 0

\

“E 10+

-

7

0

-

6a

1

0

-

~

1

1

1

1.0

1

I

0.5

I

MOLARITY OF NOCIO,

I

0.i WATER CONTENT, f,

Flgure 1. Selfdiffusion coefficients of Na’ in GTA-H20 mixtures (1.4

M NaC104, 25

(moles/kg solvent)

0.05

OC).

The salt concentration (NaC104)was 1.4 M in each solution to ensure miscibility. The effect of water content, expressed as volume fraction f,, on self-diffusion coefficients, D, of Na’ in GTA-water-NaC104 mixtures at constant salt concentration is given in Table I. The values of D decrease from 1.26 X cmz/s in the 1.4 M salt solution (no GTA) to 1.6 X lo-’ cm2/s in a 95.5% GTA. A plot of log D vs. log f , (Figure 1) is approximately a straight line over most of the composition range with slope S d = 1.5. This slope is very much smaller than the one we observed for the diffusion of sodium ions in polyethylene glycolwater mixtures’’ where S d was approximately 3.3. This is not surprising since, in the latter case, the organic solute can almost surely be considered an obstruction with a highly nonspherical shape (presumably an oblate elipsoid of revolution). The slope is also very much less than we observed’’ in a number of membranes and porous bodies where S d was about 4.6. In an earlier paper” we had shown that the log-log slope, s d , is numerically equal to the coefficient a d in the linear equation

D/Do = 1 where the “solid” fraction is 4 = (1 - f,) and Do is the diffusion coefficient a t 4 = 0. With spherical “obstructions” a d is expected to be 1.5;12our experimental value of S d thus implies that for Na’ diffusion GTA acts similar to a “dispersion” of spherical obstructions. The relative viscosity, q/qo (qo is viscosity of aqueous 1.4 M NaC104), increases from 1.00 in the aqueous salt solution to 112 in 95.5% GTA. A plot of the logarithm of the relative viscosity vs. log f , has considerable curvature at high water contents with slope S, much larger than 1.5. The plot eventually becomes a straight line with slope S, near 1.5. As a result, the relative diffusion-viscosity product Dq/D,,qo is not quite independent of solution composition but goes through a shallow maximum near f , = 0.5. The fact that the absolute value of S, is initially much larger than s d is qualitatively consistent with theory;

Flgure 2. (A) Selfdiffusion coefficients of Na’ and (B) relative viscosity for ca. 95 vol % GTA.

thus, a, in the (linear) Einstein equation ( q / q o = 1 + a,+) for spherical obstructions is 2.5. Sodium perchlorate causes a striking increase in the viscosity of GTA solutions and a significant decrease in the self-diffusion coefficient of Na’. As shown in Figure 2, viscosity (at fv = 0.03) increases by a factor of ca. 20 as the sodium perchlorate concentration increases from 0.1 to 1.6 m;in the same range, D decreases by approximately a fador of 5. While the logarithm of the viscosity increases approximately linearly with molality of NaC104, the equivalent diffusion coefficient relationship has some curvature. As a result log ( 0 7 ) increases (nonlinearly) with sodium perchlorate concentration. We report these observations here, although we do not have an explanation.

References and Notes (1) Jointly sponsored by the Offlce of Saline Water, US. Department of the Interior, and the U.S. Energy Research and Development Adminlstratlon under contract with Union Carbide Corporation. (2) Previous paper in series: R. J. Raridon and K. A. Kraus, J. Chem. €ng. Data, 16, 241 (1971). (3) Present address: EBASCO, New York, N.Y. (4) Present address: Unlon Carbide Corp., South Charleston, W.Va. (5) See, e.g., (a) K. A. Kraus, R. J. Raridon, and W. H. Baldwin, J. Am. Chem. Soc., 86, 2571 (1964); (b) J. S. Johnson, L. Dresner, and K. A. Kraus, “Hyperfiltration (Reverse Osmosis)” in “Principles of Desalination”, K. S. Spiegler, Ed., Academic Press, New York, N.Y., 1966, p 395. (6) R. J. Raridon and K. A. Kraus, J. Co//oid. Sci., 20, 1000 (1965). (7) (a) F. Nelson and K. A. Kraus, “Some Techniques for Isolating and Using Short-Lived Radioisotopes” in “Production and Use of Short-Lived Isotopes from Reactors”, Vol. I.IAEA, Vienna, 1962, p 191. (b) A. E. Marcinkowsky, F. Nelson, and K. A. Kraus, J Phys. Chem., 69 303 (1965). (8) A. E. Marcinkowsky, H. 0. Phillips, and K. A. Kraus, J. Phys. Chem, 69 3968 (1965). (9) A. E. Marcinkowsky, H. 0. Phillips, and K. A. Kraus, J. Phys. Chem, 72, 1201 (1968). (10) H. 0. Philllps, A. E. Marcinkowsky, S. B. Sachs, and K. A. Kraus, J. Phys. Chem., preceding paper in this issue. (1 1) K. A. Kraus, H. 0. Phillips, and A. E. Marcinkowsky, “Saline Water Conversion Repot?for 1965”, U. S. Department of the Interior, Office of Saline Water, p 24 and unpublished results. (12) See, e.g., (a) R. H. Stokes, “Mobilities of Ions and Uncharged Molecules in Relation to Viscoskty-A Classical Viewpoint”, in “The Structure of Electrolytic Solutions”, W. J. Hamer, Ed., Wiley, New York, N.Y., 1959, p 298. (b) J. H. Wang, J. Am. Chem. Soc., 76, 4755 (1954).

The Journal of Physical Chemistw, Vol. 81, No. 7, 1977