Comment on “Prediction of Soil Sorption Coefficients Using Model

Comment on “Prediction of Soil Sorption Coefficients Using Model Molecular Structures for Organic Matter and the Quantum Mechanical COSMO-SAC Modelâ...
0 downloads 4 Views 151KB Size
Correspondence/Rebuttal pubs.acs.org/est

Comment on “Prediction of Soil Sorption Coefficients Using Model Molecular Structures for Organic Matter and the Quantum Mechanical COSMO-SAC Model” n their article, Phillips et al.1 present what they term “[a] new method ... to predict KOC [organic carbon normalized soil−water equilibrium partition coefficient] for nonionic organic compounds that requires only molecular structures. No calibration is performed.” The authors go on to state that “[t]he experimental KOC data set ...was used to evaluate the accuracy of the KOC predictions. This data set contains KOC values for 440 nonionic organic compounds, some of which represent averages of measured KOC values from multiple soil and sediment types and different studies.” Of these 440 purportedly nonionic compounds presented in the Supporting Information of Phillips et al.,1 the following compounds have acidic and/or basic functional groups that would result in their substantial (and in many cases, effectively complete) ionization in natural surface aquatic systems and in soils and groundwaters (experimental or estimated pKa values are provided in parentheses): acetic acid (4.762); acridine (5.58 2 ); 4-aminobenzoic acid (4.78 3); aniline (4.87 2 ); anthracene-9-carboxylic acid (∼3.64); asulam (4.15); benzidine (4.652); benzo[f]quinoline (5.156), benzoic acid (4.202), 2,2′biquinoline (4.183); bromacil (9.307); 1-butylamine (10.602); chloramben (∼3.48); chlorimuro (2.033); 2-chlorophenol (8.56 2 ); 6-chloropicolinic acid (3.55 9 ); chlorsulfuron (∼3.410); 2,4-dichlorophenoxy acetic acid (2.64−3.3111); 2,5dichloro-6-methoxybenzoic acid (2.603); 3,4-dichlorophenol (8.6312); 2,3-dichlorophenol (7.7012); 3,6-dichlorosalicylic acid (1.9513); diflubenzuron (8.6012); 3,5-dinitrobenzoic acid (2.7314); 2-methyl-4,6-dinitrophenol (4.3115); hexanoic acid (4.852); 4-hydroxybenzoic acid (4.572); isocil (8.143); maleic hydrazide (5.6516), N-methylaniline (4.852), 3-methylaniline (4.712), 4-methylaniline (5.082); 3-methyl-4-bromoaniline (4.033); 1-naphthalenamine (3.922); N,N-dimethylaniline (5.072); 4-nitrobenzoic acid (3.432); pentachlorophenol (4.7417); phenylacetic acid (4.3018); phthalic acid (2.942); 4nitrophenol (7.152); pyridine (5.232); quinoline (4.902); picloram (∼2.3 1 9 ); pirimicarb (4.54 2 0 ); 2-(2,4,5trichlorophenoxy)propanoic acid (2.8417); 2,4,5-trichlorophenoxy acetic acid (2.85−3.4621); terbacil (9.017); 2,3,4,6tetrachlorophenol (5.2222); thiabendazole (4.7323); 2,4,6trichlorophenol (6.1524); 3,4,5-trichlorophenol (7.7324); 2,4,5trichlorophenol (7.0725); 3,5,6-trichloro-2-pyridinol (4.626); triclopyr (2.727); 3-trifluoromethyl-4-nitrophenol (6.0728); 3,4-dinitrobenzoic acid (2.443); dinoseb (4.6224); fenac (3.823); imazalil (6.5310); and sulfometuron methyl (3.810). Consequently, experimental KOC values for these compounds would be pH dependent, a key point that does not appear to have been considered by Phillips et al.1 Furthermore, Phillips et al.1 present a number of molecular models for terrestrial and aquatic humic acids and aquatic fulvic acids in the SMILES molecular language (see Table S1 in ref 1). The SMILES formats for these model humic and fulvic acids appear to be in the neutral forms of each compound. By

I

© XXXX American Chemical Society

definition, humic and fulvic acids are predominantly (if not nearly entirely) dissociated into their anionic forms in natural waters and moist soil systems. Basic amino moieties on some of these model humic and fulvic acids may also be protonated under environmentally relevant conditions. Consequently, it is unclear how these neutral form model humic and fulvic acid structures can be used to accurately model interactions with solutes (particularly ionizable solutes as listed above), given that in environmental systems these macromolecules would be ionized.

Sierra Rayne*



Chemologica Research, P.O. Box 74, 318 Rose Street, Mortlach, Saskatchewan S0H 3E0, Canada

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Phillips, K.; Di Toro, D.; Sandler, S. Prediction of soil sorption coefficients using model molecular structures for organic matter and the quantum mechanical COSMO-SAC model. Environ. Sci. Technol. 2011, 45, 1021−1027. (2) Haynes, W. CRC Handbook of Chemistry and Physics, 93rd ed.; Taylor and Francis: Boca Raton, FL, 2012. (3) Karickhoff, S. W.; Carreira, L. A.; Hilal, S. H. SPARC, version 4.5, September 2009. http://ibmlc2.chem.uga.edu/sparc/ (accessed May 19, 2013). (4) Hawley, J.; Bampos, N.; Abraham, R.; Sanders, J. Carboxylate and carboxylic acid recognition by tin(IV) porphyrins. Chem. Commun. 1998, 661−662. (5) Giussani, A.; Pou-Amerigo, R.; Serrano-Andres, L.; FreireCorbacho, A.; Martinez-Garcia, C.; Fernandez, M.; Sarakha, M.; Canle, M.; Santaballa, J. Combined theoretical and experimental study of the photophysics of asulam. J. Phys. Chem. A 2013, 117, 2125−2137. (6) van Vlaardingen, P.; Steinhoff, W.; de Voogt, P.; Admiraal, W. Property−toxicity relationships of azaarenes to the green alga Scenedesmus acuminatus. Environ. Toxicol. Chem. 1996, 15, 2035− 2042. (7) Bromacil; Toxicology Data Network, United States National Library of Medicine: Bethesda, MD, 2013. (8) Xu, Y.; Qin, W.; Lau, Y.; Li, S. Combination of cationic surfactant-assisted solid phase extraction with field-amplified sample stacking for highly sensitive analysis of chlorinated acid herbicides by capillary zone electrophoresis. Electrophoresis 2005, 26, 3507−3517. (9) Geronimo, J.; Smith, L.; Stockdale, G.; Goring, C. Comparative phytotoxicity of nitrapyrin and its principal metabolite, 6-chloropicolinic acid. Agron. J. 1973, 65, 689−692. (10) Tomlin, C. The e-Pesticide Manual, 13th ed., version 3.1; British Crop Protection Council: Surrey, U.K., 2004.

A

dx.doi.org/10.1021/es401135s | Environ. Sci. Technol. XXXX, XXX, XXX−XXX

Environmental Science & Technology

Correspondence/Rebuttal

(11) Environmental Health Criteria 84: 2,4-Dichlorophenoxyacetic Acid (2,4-D) - Environmental Aspects; International Programme on Chemical Safety, World Health Orgnization: Geneva, Switzerland: 1989. (12) Serjeant, E. P.; Dempsey, B. Ionisation Constants of Organic Acids in Aqueous Solution; IUPAC Chemical Data Series No. 23; International Union of Pure and Applied Chemistry (IUPAC), Pergamon Press: New York, 1979. (13) Murray, M.; Hall, J. Sorption-desorption of dicamba and 3,6dichlorosalicylic acid in soils. J. Environ. Qual. 1989, 18, 51−57. (14) Armarego, W.; Chai, C. Purification of Laboratory Chemicals; Butterworth-Heinemann: Waltham, MA, 2009. (15) Escher, B.; Schwarzenbach, R. Mechanistic studies on baseline toxicity and uncoupling of organic compounds as a basis for modeling effective membrane concentrations in aquatic organisms. Aquat. Sci. 2002, 64, 20−35. (16) Chicharro, M.; Bermejo, E.; Ongay, S.; Zapardiel, A. Determination of maleic hydrazide in potato samples using capillary electrophoresis with dual detection (UV-electrochemical). Electroanalysis 2008, 20, 534−541. (17) Mackay, D.; Shiu, W.; Ma, K.; Lee, S. Handbook of Physical− Chemical Properties and Environmental Fate for Organic Chemicals, 2nd ed.; CRC Press: Boca Raton, FL, 2010. (18) Dewick, P. Essentials of Organic Chemistry; John Wiley & Sons: New York, 2006. (19) Prevention, Pesticides and Toxic Substances; 7508W, EPA-738-F95-018; United States Environmental Protection Agency: Washington, DC, 1995. (20) Chamberlain, K.; Evans, A.; Bromilow, R. 1-Octanol/water partition coefficient (Kow) and pKa for ionisable pesticides measured by a pH-metric method. Pest. Sci. 1996, 47, 265−271. (21) Hornsby, A.; Wauchope, R.; Herner, A. Pesticide Properties in the Environment; Springer: New York, 1996. (22) Gerhartz, W. Ullmann’s Encyclopedia of Industrial Chemistry, 5th ed., Vol. A1; VCH Publishers: Deerfield Beach, FL, 1985. (23) Review Report for the Active Substance Thiabendazole; Directorate-General Health & Consumer Protection, European Commission: Brussels, Belgium, 2001. (24) Schwarzenbach, R.; Gschwend, P.; Imboden, D. Environmental Organic Chemistry; John Wiley & Sons: New York, 2005. (25) Taylor, E. Toxicology of Aquatic Pollution: Physiological, Cellular and Molecular Approaches; Cambridge University Press: Cambridge, U.K., 1996. (26) Somasundaram, L.; Coats, J.; Racke, K.; Shanbhag, V. Mobility of pesticides and their hydrolysis metabolites in soil. Environ. Toxicol. Chem. 1991, 10, 185−194. (27) Getty, D.; Getsinger, K.; Woodburn, K. A review of the aquatic environmental fate of triclopyr and its major metabolites. J. Aquat. Plant Manage. 2003, 41, 69−75. (28) Smith, M.; Applegate, V.; Johnson, B. Physical properties of some halo-nitrophenols. J. Chem. Eng. Data 1961, 6, 607−608.

B

dx.doi.org/10.1021/es401135s | Environ. Sci. Technol. XXXX, XXX, XXX−XXX