Oxidation Kinetics at Circumneutral pH - American Chemical Society

Dec 20, 2011 - Southern Cross GeoScience, Southern Cross University, Lismore, New South Wales 2480, Australia. •S Supporting Information. ABSTRACT: ...
0 downloads 0 Views 667KB Size
Article pubs.acs.org/est

Effects of pH, Chloride, and Bicarbonate on Cu(I) Oxidation Kinetics at Circumneutral pH Xiu Yuan,† A. Ninh Pham,† Guowei Xing,† Andrew L. Rose,‡ and T. David Waite*,† †

School of Civil and Environmental Engineering, The University of New South Wales, Sydney, New South Wales 2052, Australia Southern Cross GeoScience, Southern Cross University, Lismore, New South Wales 2480, Australia



S Supporting Information *

ABSTRACT: The oxidation kinetics of nanomolar concentrations of Cu(I) in NaCl solutions have been investigated over the pH range 6.5−8.0. The overall apparent oxidation rate constant was strongly affected by chloride, moderately by bicarbonate, and to a lesser extent by pH. In the absence of bicarbonate, an equilibriumbased speciation model indicated that Cu+ and CuClOH− were the most kinetically reactive species, while the contribution of other Cu(I) species to the overall oxidation rate was minor. A kinetic model based on recognized key redox reactions for these two species further indicated that oxidation of Cu(I) by oxygen and superoxide were important reactions at all pH values and chloride concentrations considered, but back reduction of Cu(II) by superoxide only became important at relatively low chloride concentrations. Bicarbonate concentrations from 2 to 5 mM substantially accelerated Cu(I) oxidation. Kinetic analysis over a range of bicarbonate concentrations revealed that this was due to formation of CuCO3−, which reacts relatively rapidly with oxygen, and not due to inhibition of the back reduction of Cu(II) by formation of Cu(II)−carbonate complexes. We conclude that the simultaneous oxygenation of Cu+, CuClOH−, and CuCO3− is the rate-limiting step in the overall oxidation of Cu(I) under these conditions.

1. INTRODUCTION Copper is an essential trace element that is vital to the viability of almost all organisms.1 In natural waters, copper typically occurs in either the cuprous (Cu(I)) or the cupric (Cu(II)) oxidation state. Copper redox chemistry in the upper water column plays a significant role in its speciation, transport, and bioavailability.2 Most previous studies3−5 have focused primarily on Cu(II), principally because Cu(I) is easily oxidized to Cu(II) by oxygen or other oxidants. However, a growing body of evidence6−11 indicates that a number of potentially important reactions, many of which are photochemically induced, may lead to Cu(I) formation and result in a significant steadystate concentration of Cu(I) in natural waters.2,12 Copper is also considered to be an important scavenger of superoxide (O2−)13,14 as it can be subjected to reduction and oxidation by O2− in natural waters.15 Zafiriou et al.16 indicated that up to 25% of inorganically complexed copper might be present as Cu(I) in sunlit seawaters provided the steady-state concentration of O2− exceeds 10−12 M. On the other hand, redox reactions of Cu(I) could result in production of reactive oxygen species (ROS), such as O2− and OH•, that may subsequently induce a cascade of radical-promoted reactions with other constituents in natural waters.17−19 As such, a better understanding of copper-catalyzed production and consumption of O2− is important in furthering insight into factors contributing to global biogeochemical cycles. © 2011 American Chemical Society

Oxygenation of micromolar and nanomolar concentrations of Cu(I) in natural seawater and synthetic NaCl solutions has been previously investigated.6,8,10,11 In general, the overall oxidation rate of Cu(I) can be described by

d[Cu(I)] = kapp[O2 ][Cu(I)] (1) dt where [Cu(I)] is the total or analytical Cu(I) concentration and kapp (in M−1·s−1) is the apparent oxidation rate constant. In oxygen-saturated systems (where [O2] ≫ [Cu(I)]), the initial decay of Cu(I) was observed to follow pseudo-first-order kinetics with a pseudo-first-order rate constant k′ = kapp[O2]0. Despite differing experimental conditions, chloride concentration has been consistently and significantly observed to influence kapp. However, large differences in calculated secondorder rate constants for individual Cu(I) species between studies suggests considerable uncertainty with regard to the actual species contributing to the overall oxidation process. For example, González-Dávila et al.6 found a significant discrepancy between the reported rate constants at micromolar and nanomolar concentrations of Cu(I) which could not be attributed entirely to the differences in the experimental conditions. More −

Received: Revised: Accepted: Published: 1527

September 27, 2011 December 16, 2011 December 20, 2011 December 20, 2011 dx.doi.org/10.1021/es203394k | Environ. Sci. Technol. 2012, 46, 1527−1535

Environmental Science & Technology

Article

for 30 min with zero-grade air that had been bubbled through 1 M NaOH to remove trace CO2 present. The pH was then quickly adjusted to specific values by adding appropriate amounts of concentrated NaOH, and the solution was subsequently continuously sparged throughout the course of each experiment. Calibration curves were developed at each pH examined by measuring the absorbance at 484 nm of 50, 100, 200, 400, and 600 nM Cu(I) added to buffer solutions containing 0.05 mM BC and 0.25 mM EDTA. The method detection limit was ∼0.5 nM Cu(I). 2.3. Speciation Modeling. Cu(I) speciation was calculated with the program MINEQL+23 using the equilibrium reactions and stability constants shown in Table 1. Extrapolation of

importantly, to the best of our knowledge, none of the previous studies6,10,11,20 has convincingly explained the variation of rate constants with varying pH and bicarbonate concentrations. The weak pH dependence of the oxidation rate constant of Cu(I) has been attributed to formation of mixed copper hydroxo− chloro complexes and back reduction of Cu(II) to Cu(I).8 The effect of bicarbonate, on the other hand, has been attributed to retardation of the back reduction of Cu(II) to Cu(I)8 and/or formation of Cu(I) carbonate complexes.11,20 As such, a clear need exists to resolve the apparent knowledge gaps in our understanding of factors controlling the kinetics of inorganic Cu(I) oxidation. In this study, oxidation of nanomolar Cu(I) was investigated over the pH range 6.5−8.0 in solutions containing different concentrations of NaCl and bicarbonate to answer two particular questions: (a) which copper species are responsible for the pH, chloride, and bicarbonate dependence of the Cu(I) oxidation rate constant and (b) how important is the back reduction of Cu(II) to the kinetics of Cu(I) oxidation? To rationalize the chemical mechanisms behind the observed results and develop a more comprehensive understanding of Cu(I) oxidation in natural waters, two different modeling approaches were adopted: (a) an equilibrium speciation modeling approach, involving assessment of the contribution of various Cu(I) species to the overall oxidation and identification of the species responsible for the pH and chloride and bicarbonate concentration dependence of Cu(I) oxidation and (b) a kinetic modeling approach where the importance of various reaction pathways was examined.

Table 1. Stability Constants for Cu(I) Speciesa no.

species

logK at 25 °C, I = 0

ref

1 2 3 4 5 6 7 8 9 10 11

H+ + OH− = H2O H+ + CO32− = HCO3− 2H+ + CO32− = H2CO3 Cu+ + Cl− = CuCl(aq) Cu+ + 2Cl− = CuCl2− Cu+ + 3Cl− = CuCl32− Cu+ + Cl− + H2O = CuClOH− + H+ Cu+ + 2Cl− + H2O = CuCl2OH2− + H+ Cu+ + Cl− + 2H2O = CuCl(OH)22− + 2H+ Cu+ + H+ + CO32− = CuHCO3 Cu+ + CO32− = CuCO3−

14.0 10.30 16.70 3.10 5.68 5.02 −5.64 −7.91 −19.84 7.0. However, in the

Figure 4. Evaluation of the deduced rate constants for oxygenation of Cu(I) species in the context of Marcus Theory. The solid line corresponds to eq 12 with λ = 84 kJ·mol−1.

presence of environmentally relevant bicarbonate concentrations, formation of CuCO3− substantially increased Cu(I) oxidation rates. Although CuCl2− and CuCl32− were the dominant Cu(I) species in this pH range in both the absence and the presence of bicarbonate, their contribution to the overall rate constant was found to be negligible. Therefore, under 1533

dx.doi.org/10.1021/es203394k | Environ. Sci. Technol. 2012, 46, 1527−1535

Environmental Science & Technology

Article

(9) Moffett, J. W.; Zika, R. G. Reaction kinetics of hydrogen peroxide with copper and iron in seawater. Environ. Sci. Technol. 1987, 21, 804− 810. (10) Sharma, V. K.; Millero, F. J. The oxidation of Cu(I) in electrolyte solutions. J. Solution Chem. 1988, 17, 581−599. (11) Sharma, V. K.; Millero, F. J. Oxidation of copper(I) in seawater. Environ. Sci. Technol. 1988, 22, 768−771. (12) Zika, R. G. In Elsevier Oceanography Series; Duursma, E. K., Dawson, R., Eds.; Elsevier: New York, 1981; Vol. 31, Chapter 10 (Marine Organic Photochemistry), pp 299−325. (13) Rose, A. L.; Waite, T. D. Kinetic model for Fe(II) oxidation in seawater in the absence and presence of natural organic matter. Environ. Sci. Technol. 2002, 36, 433−444. (14) Santana-Casiano, J. M.; González-Dávila, M.; Millero, F. J. Oxidation of nanomolar levels of Fe(II) with oxygen in natural waters. Environ. Sci. Technol. 2005, 39, 2073−2079. (15) Morel, F. M. M.; Price, N. M. The biogeochemical cycles of trace metals in the oceans. Science 2003, 300, 944−947. (16) Zafiriou, O. C.; Voelker, B. M.; Sedlak, D. L. Chemistry of the superoxide radical (O2−) in seawater: Reactions with inorganic copper complexes. J. Phys. Chem. A 1998, 102 (28), 5693−5700. (17) Goldstone, J. V.; Voelker, B. M. Chemistry of superoxide radical in seawater: CDOM associated sink of superoxide in coastal waters. Environ. Sci. Technol. 2000, 34, 1043−1048. (18) Rose, A. L.; Waite, T. D. Reduction of organically complexed ferric iron by superoxide in a simulated natural water. Environ. Sci. Technol. 2005, 39, 2645−2650. (19) Voelker, B. M.; Sedlak, D. L.; Zafiriou, O. C. Chemistry of superoxide radical in seawater: Reactions with organic Cu complexes. Environ. Sci. Technol. 2000, 34, 1036−1042. (20) Sharma, V. K.; Millero, F. J. Effect of ionic interactions on the rates of oxidation of Cu(I) with O2 in natural waters. Mar. Chem. 1988, 25, 141−161. (21) Millero, F. J.; Hershey, J. P.; Fernandez, M. The pK* of TRISH+ in Na-K-Mg-Ca-Cl-SO4 brines--pH scales. Geochim. Cosmochim. Acta 1987, 51, 707−711. (22) Moffett, J. W.; Zika, R. G.; Petasne, R. G. Evaluation of bathocuproine for the spectro-photometric determination of copper(I) in copper redox studies with applications in studies of natural waters. Anal. Chim. Acta 1985, 175, 171−179. (23) Schecher, W. D.; McAvoy, D. C. MINEQL+: A software environment for chemical equilibrium modeling. Comput. Environ. Urban Syst. 1992, 16, 65−76. (24) Buzko, V. Y.; Sukhno, I. V.; Pettit, L. D. Adjustment, estimation, and uses of equilibrium reaction constants in aqueous solution. Chem. Int. 2007, 29, 14−15. (25) Millero, F. J.; Yao, W.; Aicher, J. The speciation of Fe(II) and Fe(III) in natural waters. Mar. Chem. 1995, 50, 21−39. (26) Wang, M.; Zhang, Y.; Muhammed, M. Critical evaluation of thermodynamics of complex formation of metal ions in aqueous solutions III. The system Cu(I,II) -Cl- -e at 298.15 K. Hydrometallurgy 1997, 45, 53−72. (27) Grenthe, I.; Wanner, H. TDB-2 Guidelines for the Extrapolation to Zero Ionic Strength; minor revisions by Osthols, E.; version 6, Jan 2000 (http://www.nea.fr/html/dbtdb/guidelines/tdb2.pdf), 2000. (28) Sugasaka, K.; Fujii, A. A spectrophotometric study of copper(I) chloro-complexes in aqueous 5M Na(Cl, ClO4) solutions. Bull. Chem. Soc. Jpn. 1976, 49, 82−86. (29) Johnson, K. A.; Simpson, Z. B.; Blom, T. Global Kinetic Explorer: A new computer program for dynamic simulation and fitting of kinetic data. Anal. Biochem. 2009, 387, 20−29. (30) Mel’nichenko, N.; Koltunov, A.; Vyskrebentsev, A.; Bazhanov, A. The temperature dependence of the solubility of oxygen in sea water according to the pulsed NMR data. Russ. J. Phys. Chem. A, Focus Chem. 2008, 82, 746−752. (31) Sugasaka, K.; Fujii, A. Studies on the Preparation of Cuprous Oxide. VIII. A Spectrophotometric Study of Halogenocopper (I) Complexes in Aqueous 5 M Na(ClO4) Solutions. Bull. Chem. Soc. Jpn. 1980, 53, 2514−2519.

environmentally relevant conditions at circumneutral pH, oxidation of Cu(I) is expected to be controlled primarily by the relative proportions of Cu+, CuClOH−, and CuCO3−, which will in turn be governed largely by chloride and bicarbonate concentrations in natural waters of interest. The kinetic model developed adequately described the kinetics of Cu(I) oxidation over the range of conditions investigated in this study. In this model, oxidation of cuprous species by O2 was the most important pathway in the overall oxidation of Cu(I). While the oxidation of Cu(I) by O2− was also important under all conditions considered, back reduction of Cu(II) by O2− only became important at lower [Cl−]. Determination of values for apparent and intrinsic rate constants for oxidation of the critical Cu(I) species Cu+, CuClOH−, and CuCO3− by O2 over the pH range 6.5−8.0, whose accuracy is supported by Marcus theory, should greatly assist in understanding and predicting inorganic Cu(I) and Cu(II) transformations in natural waters. Application of the relationship determined by Marcus theory should enable further extension of these findings to systems containing other inorganic copper complexes whose stability constants and/or redox potentials are known.



ASSOCIATED CONTENT

S Supporting Information *

Additional method details, additional justification for the kinetic model, and one supplementary table and five supplementary figures with additional supporting data. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Phone: +61 2 9385 5059; fax: +61 2 9313 8341; e-mail: d.waite@ unsw.edu.au.



ACKNOWLEDGMENTS We gratefully acknowledge Christopher Miller’s assistance with the model calculations. This work was funded by the Australian Research Council Discovery Grant Scheme (DP0987188).



REFERENCES

(1) Lewis, A. G.; Cave, W. R. The biological importance of copper in oceans and estuaries. Oceanogr. Mar. Biol. Annu. Rev. 1982, 20, 471− 695. (2) Moffett, J. W.; Zika, R. G. Measurement of copper(I) in surface waters of the subtropical Atlantic and Gulf of Mexico. Geochim. Cosmochim. Acta 1988, 52, 1849−1857. (3) Buffle, J. A critical comparison of studies of complex formation between copper(II) and fulvic substances of natural waters. Anal. Chim. Acta 1980, 118, 29−44. (4) Byrne, R. H.; Miller, W. L. Copper(II) carbonate complexation in seawater. Geochim. Cosmochim. Acta 1985, 49, 1837−1844. (5) Borrok, D. M.; Nimick, D. A.; Wanty, R. B.; Ridley, W. I. Isotopic variations of dissolved copper and zinc in stream waters affected by historical mining. Geochim. Cosmochim. Acta 2008, 72, 329−344. (6) González-Dávila, M.; Santana-Casiano, J. M.; González, A. G.; Pérez, N.; Millero, F. J. Oxidation of copper(I) in seawater at nanomolar levels. Mar. Chem. 2009, 115, 118−124. (7) Millero, F. J.; Sharma, V. K.; Karn, B. The rate of reduction of copper(II) with hydrogen peroxide in seawater. Mar. Chem. 1991, 36, 71−83. (8) Moffett, J. W.; Zika, R. G. Oxidation kinetics of Cu(I) in seawater: implications for its existence in the marine environment. Mar. Chem. 1983, 13, 239−251. 1534

dx.doi.org/10.1021/es203394k | Environ. Sci. Technol. 2012, 46, 1527−1535

Environmental Science & Technology

Article

(32) Wehrli, B. Redox reactions of metal ions at mineral surfaces. Aquat. Chem. Kinet. 1990, 311−336. (33) Eberson, L. The Marcus theory of electron transfer, a sorting device for toxic compounds. Adv. Free Radic. Biol. Med. 1985, 1, 19−90. (34) Tratnyek, P. G.; Hoigné, J. Kinetics of reactions of chlorine dioxide (OCIO) in water-II. Quantitative structure-activity relationships for phenolic compounds. Water Res. 1994, 28, 57−66. (35) Stanbury, D. M. Reduction potentials involving inorganic free radicals in aqueous solution. Adv. Inorg. Chem. 1989, 33, 69−138. (36) Morel, F. M. M.; Hering, J. G. Principles and Applications of Aquatic Chemistry; Wiley: New York, 1993.

1535

dx.doi.org/10.1021/es203394k | Environ. Sci. Technol. 2012, 46, 1527−1535