You are on page 1of 7

Journal of Colloid and Interface Science 263 (2003) 454–460

www.elsevier.com/locate/jcis

Influence of solvent on the growth of ZnO nanoparticles


Zeshan Hu,a Gerko Oskam,b and Peter C. Searson a,∗
a Department of Materials Science and Engineering, Johns Hopkins University, Baltimore, MD 21218, USA
b Departamento de Física Aplicada, CINVESTAV-IPN, Mérida, Yuc. 97310, Mexico

Received 7 June 2002; accepted 20 February 2003

Abstract
We have synthesized ZnO nanoparticles by precipitation from zinc acetate in a series of n-alkanols from ethanol to 1-hexanol as a function
of temperature. In this system, nucleation and growth are relatively fast and, at longer times, the average particle size continues to increase
due to diffusion-limited coarsening. During coarsening, the particle volume increases linearly with time, in agreement with the Lifshitz–
Slyozov–Wagner (LSW) model. The coarsening rate increases with increasing temperature for all solvents and increases with alkanol chain
length. We show that the rate constant for coarsening is determined by the solvent viscosity, surface energy, and the bulk solubility of ZnO
in the solvent.
 2003 Elsevier Inc. All rights reserved.

Keywords: ZnO; Nanoparticles; Nucleation; Growth; Coarsening; Ostwald ripening; Epitaxial attachment

1. Introduction Most alcohols are dipolar, amphiprotic solvents with a di-


electric constant and viscosity that is dependent on the chain
The synthesis of ZnO quantum particles by precipitation length [16].
from alcohols results in stable colloids of nanometer-sized In this paper, we report on the influence of solvent on the
particles [1–7]. In contrast, synthesis from aqueous solutions synthesis of ZnO nanoparticles from zinc acetate at temper-
results in the formation of Zn(OH)2 [1,2] and hence ZnO can atures between 30 and 65 ◦ C. The influence of the solvent
only be obtained by using stabilizers [1] or by subsequent provides a means to achieve control over the ZnO nanopar-
hydrothermal treatment [8,9]. The nucleation process usu- ticle size and size distribution, which is essential for tailor-
ally involves the reaction between a divalent zinc salt ZnX2 ing optical, electrical, chemical, and magnetic properties of
with hydroxide ions, where X represents the anion. In solu- nanoparticles for specific applications.
tion phase synthesis, processes such as coarsening and ag-
gregation can compete with nucleation and growth in modi-
fying the particle size distribution in the system [10–12]. 2. Materials and methods
In previous work [6], we have shown that the nucleation
and growth of ZnO from 2-propanol at room temperature The ZnO colloids were prepared by precipitation from so-
is fast resulting in nanoparticles with an average radius of lution using Zn(CH3 CO2 )2 and NaOH. The overall reaction
about 1.5 nm. After the supersaturation has been depleted for the synthesis of ZnO nanoparticles from Zn(II) acetate
and growth is completed, the average particle size continues can be written as
to increase due to diffusion-limited coarsening [13,14]. The
Zn(CH3 CO2 )2 + 2NaOH
kinetics of nucleation and growth as well as processes such
as coarsening and aggregation are expected to be strongly → ZnO + 2Na(CH3 CO2 )2 + H2 O. (1)
dependent on the properties of the solvent [15–17]. Water The solvents used included ethanol (Pharmco, absolute
is a dipolar, amphiprotic solvent with a high dielectric con- ethanol), 1-propanol (J.T. Baker, reagent grade), 1-butanol
stant and, as a consequence, most salts are readily dissolved. (Alfa Aesar, reagent grade 99+%), 1-pentanol (Alfa Aesar,
reagent grade 99+%), and 1-hexanol (Alfa Aesar, reagent
* Corresponding author. grade 99+%). The solvents were used as received. The
E-mail address: searson@jhu.edu (P.C. Searson). water contents of the solvents were determined by Karl
0021-9797/03/$ – see front matter  2003 Elsevier Inc. All rights reserved.
doi:10.1016/S0021-9797(03)00205-4
Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460 455

Fisher titration (Brinkman 684 KF Coulometer): 25.2 mM


(ethanol), 18.2 mM (1-propanol), 28.5 mM (1-butanol),
32.4 mM (1-pentanol), and 14.3 mM (1-hexanol).
For a typical preparation, 1 mmol of zinc acetate di-
hydrate (Zn(CH3 CO2 )2 ·2H2 O; Aldrich, reagent grade) was
dissolved in 80 ml of solvent in a covered flask under vig-
orous stirring at 50 ◦ C. After cooling to room temperature,
8 ml of the transparent zinc salt solution was added to 64 ml
of the pure solvent. A 0.02 M NaOH (Aldrich, reagent grade)
solution was prepared by adding sodium hydroxide (Aldrich,
reagent grade) to the pure solvent in a covered flask under
vigorous stirring at 60 ◦ C. After cooling to room tempera-
ture, 8 ml of the sodium hydroxide solution was added to
20 ml of the pure solvent. The covered flasks containing
the zinc acetate solution and the sodium hydroxide solution
were heated to the growth temperature in a water bath. The
sodium hydroxide solution was then added to the zinc ac-
etate solution under vigorous stirring to give a total volume
of 100 ml with 0.1 mmol of zinc acetate and 0.16 mmol of
NaOH. From the overall reaction it follows that the synthesis
is carried out with a 25% excess of Zn(II).
Upon removal from the water bath, the colloids were cov-
ered and stored at room temperature. The colloids remained
transparent and stable for periods of up to a few months, at
which time they became translucent followed by the appear-
ance of a fine white precipitate.
Absorption spectra were obtained using a Shimadzu UV-
2101PC spectrophotometer. A slit width of 0.5 nm and a
sampling interval of 0.2 nm were used to record the spec-
tra between 275 and 400 nm. About 5-ml aliquots of the
colloid were withdrawn during particle growth at predeter-
mined time intervals and stored in an ice water bath prior to
measurement. A blank solution of the solvent was used as
reference.
The solubility of ZnO was determined by atomic ab-
sorption spectroscopy (Perkin-Elmer Analyst 100). Samples
were prepared as follows. ZnO powder (Aldrich, 99.9%)
with particle sizes smaller than 1 µm was suspended in Fig. 1. Absorbance spectra for ZnO colloids at 35 ◦ C as a function of time
2-propanol and ultrasonically agitated for 1 h. The suspen- (from left to right): ethanol (30, 60, 90, 120 min); 1-propanol (2, 4, 8, 16,
sion was then filtered over a 0.1-µm filter (Gelman Sciences) 30, 60, 93, 120 min); 1-butanol (2, 4, 8, 16, 30, 60, 90, 120 min); 1-pentanol
and the remaining solid was collected. The resulting ZnO (2, 4, 8, 16, 30, 67, 95, 120 min); 1-hexanol (2, 4, 8, 16, 30, 75, 95, 122 min).
particles with sizes larger than 0.1 µm were suspended in
2-propanol for 3 days at room temperature. The suspension exhibit a well-defined exciton peak and absorbance onset
was filtered over a 0.1-µm filter before analysis of the Zn(II) characteristic of solid ZnO. In all cases, the absorption on-
concentration with atomic absorption spectroscopy (AAS). set is significantly blue-shifted compared to the absorption
The multiple filtration steps are necessary since nanopar- onset for bulk zinc oxide at about 385 nm, showing that the
ticles are also detected by AAS, which would result in an average particle size is in the quantum regime. For all sol-
overestimation of the ZnO solubility. Calibration was per- vents, the absorption onset red-shifts with time due to the
formed by preparing solutions of Zn(ClO4 )2 of known con- increase in average particle size.
centration in the mixed solvents. The influence of the solvent can be clearly seen during the
nucleation process after mixing the zinc acetate and sodium
3. Results and discussion hydroxide solutions. For the longer chain length alcohols,
the spectra show a well-defined absorbance onset imme-
Figure 1 shows absorption spectra for the synthesis of diately after mixing, indicating that nucleation is fast. In
ZnO from zinc acetate in a series of alcohols from ethanol contrast, for the shorter chain length alcohols, ethanol and
to 1-hexanol at 35 ◦ C. In general, the absorbance spectra 1-propanol, the absorbance spectra evolve with time, indi-
456 Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460

Fig. 3. Transmission electron microscope image of ZnO particles grown


from Zn(CH3 COO)2 in propanol at 55 ◦ C for 8.5 h.

termine an average diameter of 6.5 ± 1.2 nm (289 particles).


The average particle diameter obtained from the absorption
onset is 6.9 nm, in good agreement with the value obtained
from analysis of transmission electron microscope images.
We note, however, that the particle diameter is in the range
Fig. 2. Band gap, E ∗ , and absorption onset wavelength, λ, for ZnO quan-
tum particles versus particle radius, calculated from Eq. (1), using me =
where the band gap enlargement is small and hence the error
0.26, mh = 0.59, and ε = 8.5. The onset wavelength was calculated from in determining the particle size is relatively large.
E ∗ = hc/λ. Figure 4 shows the time dependence of the average parti-
cle radius, obtained from the absorption onset and Eq. (2), at
cating that nucleation and growth are slower. The accelera- different temperatures. In all solvents, the average radius in-
tion of the nucleation and growth with decreasing dielectric creases with time. In addition, at any time it is seen that the
constant has been reported previously for ZrO2 particles in particle size increases with increasing temperature. For all
mixed aqueous/alcohol solution [18,19]. solvents the initial particle radius is between 1.2 and 1.5 nm
The average particle size in a colloid can be obtained depending on temperature, indicating that the mechanism for
from the absorption onset using the effective mass model nucleation is similar in all cases.
(e.g., [20]) where the band gap E ∗ (in eV) can be approxi- Coarsening involves the growth of larger crystals at the
mated by expense of smaller crystals and is governed by capillary ef-
  fects. Since the chemical potential of a particle increases
h̄2 π 2 1 1 1.8e with decreasing particle size, the equilibrium solute concen-
E ∗ = Egbulk + 2 ∗
+ ∗ −
2er me m0 mh m0 4πεε0r tration for a small particle is much higher than that for a
3  −1 large particle. The resulting concentration gradients lead to
0.124e 1 1
− 2 ∗
+ ∗ , (2) the transport of solute (e.g., metal ions) from the small par-
2
h̄ (4πεε0 ) me m0 mh m0 ticles to the larger particles. The rate law for this process,
where Egbulk is the bulk band gap (eV), h̄ is Plank’s constant, derived by Lifshitz, Slyozov, and Wagner (LSW) [23,24], is
r is the particle radius, me is the electron effective mass, mh given by
is the hole effective mass, m0 is free electron mass, e is the
r̄ 3 − r̄03 = kt, (3)
charge on the electron, ε is the relative permittivity, and ε0
is the permittivity of free space. Figure 2 shows the band where r̄ is the average particle radius, r̄0 is the average initial
gap and corresponding absorption onset wavelength versus particle radius, k is the rate constant, and t is time. Figure 5
particle radius calculated using Eq. (2). Due to the rela- shows the growth data from Fig. 3 re-plotted as r̄ 3 versus
tively small effective masses for ZnO (me = 0.26, mh = 0.59 time. The linear region at longer times is consistent with
[21,22]), band gap enlargement is expected for particle radii rate law for coarsening and indicates that particle growth is
less than about 4 nm. completed shortly after nucleation so that at longer times the
Figure 3 shows a transmission electron microscope im- increase in particle size is determined solely by diffusion-
age of ZnO particles grown from Zn(CH3 COO)2 at 55 ◦ C limited coarsening. Since the slope of the linear region of
after 8.5 h. From the lattice fringes it is seen that the par- the curves corresponds to the rate constant for coarsening, it
ticles are approximately spherical and that many particles is evident that the rate constant is dependent on the temper-
exhibit some faceting. From analysis of many images we de- ature and solvent.
Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460 457

Fig. 4. ZnO particle radius versus time determined from the results in Fig. 1 Fig. 5. The results from Fig. 3 re-plotted as r 3 versus time at (4) 75 ◦ C,
and Eq. (1) for different temperatures: (4) 75 ◦ C, (!) 65 ◦ C, (1) 55 ◦ C, (!) 65 ◦ C, (1) 55 ◦ C, (e) 45 ◦ C, (P) 35 ◦ C, (E) 21 ◦ C, and (%) 0 ◦ C.
(e) 45 ◦ C, (P) 35 ◦ C, (E) 21 ◦ C, and (%) 0 ◦ C.

temperature and solvent dependence of the rate constant can


Figure 6 shows the rate constant k versus 1/T for growth
be considered by eliminating the diffusion coefficient from
from the series of alcohols determined from Fig. 5. The rate
Eq. (4) using the Stokes–Einstein equation,
constant is strongly temperature dependent and increases
with increasing alcohol chain length. The rate constant k is kB T
D= , (5)
given by [6] 6πηa
8γ DVm2 cr=∞ where kB is the Boltzmann constant, η is the viscosity of the
k= , (4) solvent, and a is the solvated ion radius. Substituting into
9RT
Eq. (4), we obtain
where γ is the surface energy, D is the diffusion coeffi-
cient, Vm is the molar volume, and cr=∞ is the equilibrium 8γ Vm2 cr=∞
concentration at a flat surface (i.e., the bulk solubility). The k= . (6)
54πηaNA
458 Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460

Fig. 7. Viscosity versus temperature for (a) ethanol, (b) 1-propanol, (c) 1-bu-
tanol, (d) 1-pentanol, and (e) 1-hexanol. Data obtained from Ref. [33].

Fig. 6. The rate constant for coarsening determined from the slopes of
the r 3 versus time curves versus inverse temperature for (%) ethanol,
(1) 1-propanol, (P) 1-butanol, (!) 1-pentanol, and (e) 1-hexanol.

Thus, the rate constant for coarsening is a function of


γ (the surface energy of the crystal), Vm (the molar vol-
ume), cr=∞ (the bulk solubility), η (the solvent viscosity),
and a (the solvated ion radius). The molar volume for ZnO
is 14.8 cm3 mol−1 and is essentially constant since the lin-
ear expansion coefficient is about 4 × 10−6 K−1 over a wide
temperature range [22]. The surface energy for the solid–
vapor interface for metal oxides is on the order of 1 J m−2
[25–27]; however, adsorption of ions and solvent molecules
is expected to reduce the surface energy of the solid–liquid
interface to values in the range 0.1–0.5 J m−2 [28]. In the
gas phase, methanol, ethanol, and propanol adsorb dissocia- Fig. 8. The rate constant for coarsening versus the viscosity of the sol-
tively on ZnO under ambient conditions [29–31]; however, vent: (%) ethanol, (1) 1-propanol, (P) 1-butanol, (!) 1-pentanol, and
(e) 1-hexanol. The solid lines connect the data points for the same solvent,
data for the surface energy of solids in alcohols are not avail-
while the dotted lines connect the data points for constant temperature.
able in the literature.
The solvated ion radius for Zn(II) in methanol is 0.51 nm
and is independent of temperature over the temperature perature, the rate constant increases with solvent viscosity
range of interest [32]. The ionic radius for Zn(II) is 0.075 nm (solvent chain length). From Eq. (6) it is apparent that the
and the length of a methanol molecule is about 0.36 nm. rate constant k ∝ η−1 if cr=∞ and γ are independent of the
Thus, the solvated ion radius implies a high coordination solvent. This dependence is related to the fact that the dif-
number of the solvent molecules. Since 1-hexanol is about fusion coefficient for zinc ions in the solvent decreases with
twice as long as an ethanol molecule, the solvated ion ra- increasing viscosity. However, Fig. 8 shows that the rate con-
dius for zinc ions in hexanol would be expected to be about stant, at constant temperature, increases with increasing vis-
twice the solvated ion radius in ethanol, assuming similar cosity according to a power law relationship with exponents
coordination numbers. From Eq. (6) we see that k ∝ a −1 , so increasing from 0.21 at 0 ◦ C to 0.79 at 75 ◦ C. The positive
that an increase in the solvated ion radius due to increasing exponents in Fig. 8 imply that either the surface energy γ or
chain length would be expected to result in a decrease in the the bulk solubility cr=∞ must increase with increasing vis-
rate constant by a factor of about 2. However, from Fig. 6 cosity (solvent chain length).
it can be seen that the rate constant increases with increas- For each solvent, the rate constant decreases with increas-
ing chain length, indicating that the influence of the solvent ing viscosity (decreasing temperature) with an exponent that
on the growth kinetics is not dominated by the solvated ion decreases from −2.1 for ethanol to −1.33 for hexanol. For
radius. Hence, the dependence of the rate constant on the a given solvent, an exponent of −1 is predicted from Eq. (6)
chain length of the solvent must be dominated by the sur- if cr=∞ is independent of temperature (Vm , γ , and a are ex-
face energy γ , the solvent viscosity η, or the bulk solubility pected to be weakly dependent on temperature). Thus, the
cr=∞ . magnitude of the exponents indicate that, for a given sol-
Figure 7 shows that the viscosity of the alcohols is vent, cr=∞ increases with decreasing viscosity (increasing
strongly dependent on temperature and chain length, espe- temperature), as would be expected for most solids.
cially at lower temperatures [33]. Figure 8 shows the rate The bulk solubility of ZnO is expected to be related to
constant, k, plotted versus the solvent viscosity. At each tem- the enthalpy of solution, which contains contributions from
Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460 459

Fig. 9. The solubility of ZnO in mixtures of 2-propanol and water deter-


mined versus inverse dielectric constant of the solvent mixture.
Fig. 10. The rate constant for coarsening multiplied with the solvent
viscosity for ZnO particle growth versus inverse dielectric constant for
the lattice enthalpy and the ion solvation enthalpy. Since the (%) ethanol, (1) 1-propanol, (P) 1-butanol, (!) 1-pentanol, and (e) 1-hex-
lattice enthalpy of ZnO is independent of solvent, the influ- anol.
ence of the solvent on the bulk solubility is expected to be
through the enthalpy of solution. The bulk solubility cr=∞
versal curve except at high viscosity. These results illustrate
can be estimated from the coarsening rate constants (Fig. 6)
that the rate constant multiplied by the solvent viscosity
and Eq. (6). Taking γ = 0.1 J m−2 , Vm = 14.8 cm3 mol−1 ,
is uniquely described as a function of the dielectric con-
a = 0.5–1 nm, η = 0.01–0.1 P, and k = 10−4 –10−3 nm3 s−1 ,
stant.
the bulk solubility is determined to be in the range 10−11 –
From Eq. (6) it can be seen that kη is proportional to the
10−9 mol l−1 , and as a consequence cannot be measured
directly. product γ cr=∞ assuming that Vm and a are constant. Ac-
To determine the dependence of the solubility of ZnO on cording to Eq. (6), the increase of kη with inverse dielectric
the dielectric constant of the solvent, we measured the sol- constant implies that γ cr=∞ increases with decreasing ε.
ubility of ZnO in water/2-propanol mixtures as a function Since cr=∞ is expected to increase with increasing dielectric
of the concentration of water. These equilibrium solubil- constant, as described above, the results in Fig. 10 suggest
ity experiments were performed in solvent mixtures, thus that the increase in kη with increasing ε−1 is dominated
minimizing the influence of ion adsorption. Figure 9 shows by an increase in surface energy with increasing ε (i.e., the
the zinc ion concentration versus water concentration in surface energy increases with increasing chain length of the
water/2-propanol mixtures. Since the dielectric constant in- solvent).
creases with increasing concentration of water, Fig. 9 shows These results show that particle growth and coarsening
that the solubility of the ZnO particles increases with in- are strongly dependent on solvent through the viscosity, bulk
creasing dielectric constant ε. This dependence indicates solubility, and surface energy. The solvent is an important
that the partially covalent nature of the bonding in ZnO parameter in determining the coarsening kinetics, and that by
plays an important role in determining the dissolution mech- choice of solvent, the ZnO particle synthesis can be tuned.
anism since, for a purely ionic solid, the solubility would be
expected to increase with increasing solvent dielectric con-
stant. In pure water, the measured Zn(II) concentration is 4. Summary
49 µM, in good agreement with values reported in the lit-
erature [34]. For water concentrations less than 60 vol%, The synthesis of ZnO quantum particles by precipitation
the Zn(II) concentration is below the detection limit for from a series of n-alkanols from ethanol to 1-hexanol alco-
AAS (≈1 µM). Estimation of the solubility of Zn(II) in pure hols results in stable colloids of nanometer-sized particles.
propanol by extrapolation of the solubility data to zero water For ethanol and 1-propanol, nucleation and growth are re-
content results in values several orders of magnitude lower tarded compared to longer chain length alcohols where nu-
than that in pure water [35], consistent with the values of cleation and growth are fast. After the supersaturation has
10−11 –10−9 mol l−1 obtained from the rate constant using been depleted and nucleation and growth are completed, the
Eq. (6). average particle size continues to increase due to diffusion-
Figure 10 shows the growth rate constant multiplied by limited coarsening. For all alcohols, the kinetics of coars-
the solvent viscosity, kη, versus ε−1 . It can be seen that kη ening were consistent with the Lifshitz–Slyozov–Wagner
increases with increasing ε−1 in a similar manner for all al- model. The coarsening rate constant increases with increas-
cohols. At low ε−1 (i.e., high ε) kη increases strongly with ing temperature and longer solvent chain length, due to the
ε−1 , while at higher ε−1 (i.e., lower ε) kη is essentially in- influence of solvent viscosity, surface energy, and the bulk
dependent of ε−1 . Figure 10 shows that the kη versus ε−1 solubility of ZnO. These results illustrate that the solvent is
curves for all solvents and temperatures collapse onto a uni- an important parameter in controlling particle size.
460 Z. Hu et al. / Journal of Colloid and Interface Science 263 (2003) 454–460

Acknowledgments [17] J. Barthel, H.-J. Gores, in: G. Mamantov, A.I. Popov (Eds.), Chemistry
of Nonaqueous Solutions, VCH, New York, 1994, pp. 1–112.
[18] M.Z.-C. Hu, M.T. Harris, C.H. Byers, J. Colloid Interface Sci. 198
The authors gratefully acknowledge support from the
(1998) 87.
JHU MRSEC (NSF Grant No. DMR00-80031). G.O. ac- [19] M.Z.-C. Hu, E.A. Payzant, C.H. Byers, J. Colloid Interface Sci. 222
knowledges support from CONACYT (Grant No. 33171-E (2000) 20.
(G.O.)). [20] L.E. Brus, J. Phys. Chem. 90 (1986) 2555.
[21] S. Shionoya, in: S. Shionoya, W.M. Yen (Eds.), Phosphor Handbook,
CRC, Boca Raton, FL, 1998.
[22] L.I. Berger, Semiconductor Materials, CRC, Boca Raton, FL, 1997.
References [23] I.M. Lifshitz, V.V. Slyozov, J. Phys. Chem. Solids 19 (1961) 35.
[24] C. Wagner, Z. Elektrochem. 65 (1961) 581.
[1] U. Koch, A. Fojtik, H. Weller, A. Henglein, Chem. Phys. Lett. 122 [25] X.G. Wang, W. Weiss, Sh.K. Shaikhutdinov, M. Riter, M. Petersen,
(1985) 507. F. Wagner, R. Schlogl, M. Scheffler, Phys. Rev. Lett. 81 (1998) 1038.
[2] D.W. Bahnemann, C. Kormann, R. Hoffmann, J. Phys. Chem. 91 [26] I. Manassidis, A. De Vita, M.J. Gillan, Surf. Sci. Lett. 285 (1993)
(1987) 3789. L517.
[3] M. Haase, H. Weller, A. Henglein, J. Phys. Chem. 92 (1988) 482. [27] D.A. Weirauch, P.D. Ownby, J. Adhesion Sci. Technol. 13 (1999)
[4] L. Spanhel, M.A. Anderson, J. Am. Chem. Soc. 113 (1991) 2826. 1321.
[5] P.V. Kamat, B. Patrick, J. Phys. Chem. 96 (1992) 6829. [28] A. Zangwill, Physics at Surfaces, Cambridge University Press, Cam-
[6] E.M. Wong, J.E. Bonevich, P.C. Searson, J. Phys. Chem. B 102 (1998) bridge, 1988.
7770. [29] J.M. Vohs, M.A. Barteau, Surf. Sci. 201 (1988) 481.
[7] L. Guo, S. Yang, C. Yang, P. Yu, J. Wang, W. Ge, G.K.L. Wong, Appl. [30] J.M. Vohs, M.A. Barteau, Surf. Sci. 221 (1989) 590.
Phys. Lett. 76 (2000) 2901. [31] J.M. Vohs, M.A. Barteau, J. Phys. Chem. 95 (1991) 297.29.
[8] C.-H. Lu, C.-H. Yeh, Ceram. Int. 26 (2000) 351. [32] R. Lovas, G. Macri, S. Petrucci, J. Am. Chem. Soc. 92 (1970) 6502.
[9] D. Chen, X. Jiao, G. Cheng, Solid State Commun. 113 (2000) 363. [33] J. Barthel, R. Neueder, R. Meier (Eds.), Electrolyte Data Collec-
[10] V.K. LaMer, R.H. Dinegar, J. Am. Chem. Soc. 72 (1950) 4847. tion, Part 3: Viscosity of Nonaqueous Solutions I: Alcohol Solutions,
[11] J. Park, V. Privman, E. Matijevic, J. Phys. Chem. B 105 (2001) 11630. Chemistry Data Series, Vol. XII, Dechema, Frankfurt, 1997, Part 3.
[12] R.L. Penn, G. Oskam, T.J. Strathmann, P.C. Searson, A.T. Stone, D.R. [34] H. Remy, Z. Elektrochem. Angew. Phys. Chem. 31 (1925) 88.
Veblen, J. Phys. Chem. B 105 (2001) 2177. [35] Extrapolation of an exponential fit to the experimental data gives a
[13] Z. Hu, G. Oskam, R.L. Penn, N. Pesika, P.C. Searson, J. Phys. Chem. solubility for Zn(II) in pure propanol of 3.5 × 10−7 M. From Fig. 10
B, submitted. it is clear that the solubility decreases more rapidly than predicted by
[14] G. Oskam, Z. Hu, R.L. Penn, P.C. Searson, Phys. Rev. E, submitted. the exponential fit so that the extrapolation represents an upper limit.
[15] O. Popovich, R.P.T. Tomkins, Nonaqueous Solution Chemistry, Wiley, Since there are no suitable models that we can use to fit the data, we
New York, 1981. can only conclude that these data are consistent with the solubilities
[16] C. Kalidas, G. Hefter, Y. Marcus, Chem. Rev. 100 (2000) 819. determined from the rate constants.

You might also like