Degradation of Organophosphate Pesticides Using Pyridinium Based

Oct 7, 2016 - Center for Basic and Advanced Research, Faculty of Informatics and Management, University of Hradec Kralove, 500 03 Hradec Kralove, ...
1 downloads 0 Views 748KB Size
Subscriber access provided by UNIV NEW ORLEANS

Article

Degradation of Organophosphate Pesticides Using Pyridinium Based Functional Surfactants Rahul Sharma, Bhanushree Gupta, Toshikee Yadav, Srishti Sinha, Arvind Kumar Sahu, Yevgen Karpichev, Nicholas Gathergood, Jan Marek, Kamil Kuca, and Kallol Kumar Ghosh ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.6b01878 • Publication Date (Web): 07 Oct 2016 Downloaded from http://pubs.acs.org on October 10, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Degradation of Organophosphate Pesticides Using Pyridinium Based Functional Surfactants Rahul Sharmaa, b, Bhanushree Guptaa, Toshikee Yadava, Srishti Sinhaa, Arvind Kumar Sahua, Yevgen Karpichevc ,d *, Nicholas Gathergood c, Jan Mareke ,f,g, Kamil Kuca d ,f, Kallol K. Ghosh a* a

School of Studies in Chemistry, Pt. Ravishankar Shukla University, Raipur (C.G.) 492010, India b Amity School of Engineering & Technology, Amity University Madhya Pradesh, Gwalior, India 474005 c Chair of Green Chemistry, Department of Chemistry, Tallinn University of Technology, 12618 Tallinn, Estonia d Center for Basic and Advanced Research, Faculty of Informatics and Management, University of Hradec Kralove, 500 03 Hradec Kralove, Czech Republic e Faculty of Military Health Sciences, University of Defense, 500 01 Hradec Kralove Czech Republic f Biomedical Research Center, University Hospital Hradec Kralove, 500 05 Hradec Kralove, Czech Republic g Department of Surgical Studies, Faculty of Medicine, University of Ostrava, 700 30 Ostrava, Czech Republic

Corresponding Authors *Dr. Yevgen Karpichev, phone: +372 620-4382; fax: +372 620 2994 E-mail: [email protected] *Prof. Kallol K. Ghosh, phone: +91-771-2263146, fax: +91-771-2262583 Email: [email protected]

Keywords: Sustainable agriculture, Organophosphorus pesticides; α-Nucleophile, Oximes; Functionalized surfactants; Micellar effects

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 39

Degradation of Organophosphate Pesticides Using Pyridinium Based Functional Surfactants

Rahul Sharma, Bhanushree Gupta, Toshikee Yadav, Srishti Sinha, Arvind Kumar Sahu, Yevgen Karpichev*, Nicholas Gathergood, Jan Marek, Kamil Kuca, Kallol K. Ghosh*

Synopsis: Environmentally friendly decomposition of organophosphorus pesticides occurs in micellar solution of functionalized pyridinium surfactants

Table of Contents Graphic

2 ACS Paragon Plus Environment

Page 3 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

ABSTRACТ The enhancement of environmental quality is one of the key principles of sustainable agriculture which points lesser use of synthetic pesticides and chemical fertilizers. Green chemistry offers an array of innovative approaches to develop safe and efficient methods of its chemical transformation towards nontoxic readily biodegradable products under mild conditions. The development of new strategies for chemical decontamination of organophosphorus nerve agents and pesticides is an issue of immediate concern. Oximes have been demonstrated to find an application as functionalized organized molecular systems. In this study, kinetic investigations have been explored to estimate the nucleophilic efficiency of oxime-functionalized pyridinium surfactants 3hydroxyiminomethyl-1-alkylpyridinium bromide (alkyl=CnH2n+1, n=10, 12, 14, 16) and 4hydroxyiminomethyl-1-alkylpyridinium bromide (alkyl=CnH2n+1, n=10, 12) for the hydrolysis of pesticides paraoxon (NPDEP) and methyl paraoxon (NPDMP) in the mixed micelles with conventional cationic surfactants CPB, CTAB and CDMEAB. Comprehensive study of surface properties and acid-base equilibria of micellar system composed by (i) functionalized surfactants and (ii) mixed functionalized / conventional cationic micelles have been carried out. pKa of studied nucleophiles in the presence of surfactants has also been monitored. Effect of pH, co-micellar effect of other surfactants and alkyl chain length of functionalized surfactants has been monitored on the observed rate constants of cleavage of studied organophosphates.

3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 39

INTRODUCTION The concept of sustainable development is in consonance with safe environment and environment protection due to harmonization of all the land users guarantees safe protected habitat for human and other living organisms. Since the Rio Declaration on Environment and Development 1992 it has been pointing out that a crucial step to ensure development is to remove all negative factors leading to unsustainability.1 Enhance environmental quality is one of the key principles of sustainable agriculture and follows lesser use of synthetic pesticides and chemical fertilizers. Green chemistry offers an array of innovative approaches, safe and efficient methods of its chemical transformation towards nontoxic readily biodegradable products under mild conditions. Organophosphate esters are frequently used as pesticides, insecticides, rodenticides and other bioactive agents.2 Due to the biological significance of these esters their environmental implication and degradation have been extensively studied over the past decades. Exposure to even small amount of organophosphorus compounds is fatal and may result into death due to respiratory failure. Despite the frequent efforts to limit the misuse and mishandling of toxic organophosphorus based pesticides their toxicity resulted into 200,000 fatalities annually.3-4 The inactivation of acetylcholinesterase (AChE) leads to improper functioning of nervous system and finally to death.56

Using organophosphorus (OP) compounds as pesticides and chemical warfare agents may lead to

massive intoxication from suicidal and accidental events, or terrorist attack. In 1995, a religious sect Aum Shinrikyo used nerve agent GB (sarin) to poison people on a Tokyo subway.7 Sarin delivered by rockets was used in the chemical warfare attack in Syria in 2013.8 Chemical weapons still pose a serious concern in the age of terrorist activity and a risk of using it by rogue states. There is no disagreement among chemists regarding whether a plant which produces OP pesticides can also make nerve agents. Another example of misusing OP pesticides is Project Coast program launched in South Africa during apartheid. It was aimed to develop a range of chemical and

4 ACS Paragon Plus Environment

Page 5 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

biological agents, including pesticide paraoxon (NPDEP) being used as an assassination weapon in South Africa.9 It is noteworthy that paraoxon is one of the most potent AChE-inhibiting insecticides available. In spite of being used as an ophthalmological medicine, it is now rarely used as an insecticide due to the risk of poisoning to humans and animals. The name paraoxon is applied to two compounds being 4-nitrophenyl ester of dimethyl- (NPDMP, or methyl paraoxon) and diethyl(NPDEP, or ethyl paraoxon) phosphoric acid, the latter is usually called simply paraoxon, see Figure 1. In spite of the chemical similarity, the methyl and ethyl derivatives have been reported to demonstrate methyl vs ethyl group selectivity in some cases, e.g. to the sunfish10 whereas it has not been reported for mammals, e.g. mouse11and human.12

Figure 1. Examples of organophosphate based pesticides The facile hydrolysis of organophosphorus compounds (OP) is of theoretical and practical interest. Over the last few decades a number of different approaches have been used to address the issue of detoxification/hydrolysis of such OP compounds. These includes the reaction of these esters with some potent nucleophiles, e.g. oximes,13-14 hydroxamates,15-18 peroxides,19 oiodosocarboxylates,20 etc. The therapeutic ability of pyridinium based oximes as antidotes has been illustrated under physiological conditions and in several cases of patients intoxicated with OP insecticides.21 Numerous steps have been undertaken to develop a better nucleophile by the insertion

5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 39

of a nucleophilic moiety into the polar head group of surfactants to generate more efficient “functionalized surfactants” with high nucleophilicity and solubilizing power.22-24 Functionalized surfactants belong to the class of most potent reactants in their ability to cleave P-O bond in OP esters.25-29 From the last several years, ionic liquids (ILs) have grasped huge and ever-growing attention from both the scientific and industrial communities30 in various fields, including their use in effective and selective extraction.31 They are also utilized in designing of composite systems for detection32of OP pesticides. Among the library of ILs, few ILs with long-chain alkyl groups are capable of self-assembling to form aggregates in aqueous solutions were studied meticulously because of their inherent amphiphilic nature and nowadays may be considered as promising environment friendly solvents for solubilisation of sparingly soluble pesticides.33-34 Recently reported progress in the development of IL-based micellar-catalytic systems35 and versatile “benign by design” approach toward synthesis of biodegradable pyridinium ILs

36-39

opens wide

opportunities for using sustainable microorganized ILs systems for pesticide degradation. In spite of huge efforts devoted to design in the past several decades, synthesis and development of novel pyridinium oximes as potential antidotes for OP-intoxication, there was no fruitful advances in designing compounds effective against all kinds of nerve agents and other OPs. From the last few years, we have been involved in studying the catalytic potencies of various αnucleophiles against hydrolysis of various esters in self-organized systems.40-45 Earlier we reported the cleavage of phosphate and sulfonate (p-nitrophenyl toluene sulfonate; PNPTS) esters with pyridinium oxime based 3-series (3-Cn) and 4-series (4-Cn) functionalized surfactants.41 Effect of CPB was also studied on the rate of esterolytic reactions. In continuation herein, we have examined the micellar properties, acid dissociation constants, and hydrolytic efficacy of 3hydroxyiminomethyl-1-alkylpyridinium bromide (alkyl chain length hexadecyl, 3-C16; tetradecyl,

6 ACS Paragon Plus Environment

Page 7 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

3-C14; dodecyl, 3-C12; and decyl, 3-C10) and 4-hydroxyiminomethyl-1-alkylpyridinium bromide (alkyl chain length dodecyl, 4-C12; and decyl, 4-C10) toward the cleavage of phosphate esters NPDEP and NPDMP in mixed micelles of some cationic surfactants; cetylpyridinium bromide (CPB), cetyltrimethylammonium bromide (CTAB) and cetyl dimethyl ethyl ammonium bromide (CDMEAB) in aqueous medium (Figure 2).

Figure 2. Structures of investigated nucleophiles and conventional surfactants EXPERIMENTAL Materials Diethyl-p-nitrophenyl phosphate (NPDEP, paraoxon) and dimethyl-p-nitrophenyl phosphate (NPDMP, methyl paraoxon) were procured from Sigma Aldrich, Bangalore, India. 3- and 4hydroxyiminomethyl-1-alkylpyridinium bromide series of functionalized surfactants were prepared by quaternization of pyridine aldoxime with the corresponding alkyl bromides.46 CPB, CTAB was purchased from Sigma Aldrich Chemical Pvt. Limited, Bangalore. CDMEAB was obtained from

7 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 39

the lab of Prof. R. M. Palepu, St. Francis Xavier University, Antigonish, Canada. All solutions were prepared in triple distilled water. Methods Surface Tension Measurements The surface tension of investigated surfactants at pH 7.0 was determined with a surface tensiometer (Jencon, India) using a platinum ring by the ring detachment technique. The tensiometer was calibrated against distilled water. Platinum ring was thoroughly cleaned and dried before each measurement. The ring was hung to the balance, dipped into the solution and then pulled out. The maximum force needed to pull the ring through the interface was measured and correlated to the surface tension. Each experiment was repeated several times until good reproducibility was achieved. The results were accurate within ± 0.1 mN·m-1. Determination of Acid Dissociation Constant The acid dissociation constants (pKa) of all the functionalized surfactants (Figure 3) were determined using the method of Albert and Sergeant.47 pKa is based on direct estimation of the ratio of molecular species (protonated) to the ionized (deprotonated) species in a series of non-absorbing phosphate buffer solutions. For this purpose, the spectra of molecular species were obtained first in buffer solution of particular pH in which compounds of interest would be present wholly in either form. An aliquot (1-3 ml) of oxime stock solutions (0.5 mM) were diluted to 25 ml buffer solution (combination of 70 mM disodium hydrogen phosphate and 70 mM potassium dihydrogen phosphate), Different pH values ranging from 6.35 to 10.32 were maintained using different composition of phosphate buffer, further pH was adjusted to the desired value by the addition of dilute sodium hydroxide solution (100 mM). After each pH modification, the solution was transferred into the cuvette, and the absorption spectra were recorded. All absorbances were recorded over the wave-length range of 200–500 nm with a reference to blank solution at 27± 0.5˚C. pK a = pHexp - log Abs  AbsHOx AbsOx  Abs

(1)

8 ACS Paragon Plus Environment

Page 9 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Where, AbsΨ represents the absorbance of partially ionized oxime species at particular pH; AbsHOx is the absorbance of unionized form of oxime; and AbsOx is the absorbance of the completely ionized form of oxime.

Figure 3. Absorbance spectra of 3-hydroxyiminomethyl-1-decylpyridiniumbromide (3-C10) in 0.5 mM CPB at 27˚C. The average value of the multiple measurements was considered the pKa of the oxime functionality. Varian Cary 50 UV-Visible double beam spectrophotometer with quartz cells of 10 mm was used for spectrometric analysis. The quartz cells were attached to a thermostatic peltier for maintaining the constant temperature (27 ± 0.5 ˚C). pH values of buffers were determined using a Eutech (pH 700), pH meter equipped with Inlab@ Expert Pro glass electrode with an accuracy of ±0.01 units. The pH meter was calibrated at 27˚C using the two-point calibration method with commercially available standard buffer solutions at pH 7.00 and 9.20. The spectrophotometric determination of pKa of 3-C10 is reproduced in Figure 3 at 27˚C. Kinetic Measurements The rate of nucleophilic reaction (scheme 1) was studied at 27˚C by monitoring the increase in absorption of p-nitrophenoxide anion (400 nm) using a Varian Cary 50

9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 39

UV− visible spectrophotometer equipped with a Peltier temperature controller unit. The kinetic study was performed under pseudo-first-order conditions with the concentration of nucleophile in excess over the substrate concentration (i.e. large excess of oximate anions over the substrate (1:10) and at an ionic strength of 0.1 M (with KCl). The initial concentration of the substrate was 0.5 mM for all reactions. Borate buffer was employed to control the pH of the media. The pH of the reaction medium was measured using a Eutech pH 700, pH meter equipped with Inlab@ Expert Pro glass electrode with an accuracy of ±0.01 units. p-Nitrophenoxide ion was liberated quantitatively and identified as one of the products by comparing the UV− visible spectrum at the end of the reaction with the authentic sample under the same experimental condition. Each experiment was repeated twice, and the observed rate constants were reproducible within a precision of greater than 2%. For all the kinetic runs, the absorbance/time results fit very well to the first-order rate eq. 2: ln(A∞ - At) = ln(A∞ - A0) – kt

(2)

Scheme 1. Nucleophilic attack of oximate ion at P=O phosphate ester. The pseudo-first order rate constants (kobs) were determined for all reactions (Scheme1) from the plots of versus time with A0, At, and A∞ being the absorbance values at zero, time, and infinite time respectively. A representative graph for the hydrolysis of NPDEP with 3-C12 at 0.5 mM CPB showing an increase in absorbance of p-nitrophenoxide ion at 412 nm is presented in Figure 4.

10 ACS Paragon Plus Environment

Page 11 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 4. UV spectra collected at different reaction times showing an increase in absorbance of pnitrophenoxide ion for the cleavage of NPDEP with 3-C12 in the presence of CPB surfactant. Reaction conditions: [POX-E] = 0.05 mM, [3-C12] = 0.5 mM, [KCl] = 0.1 M, [CPB] = 0.5 mM, pH = 9.0; 27˚C. RESULTS AND DISCUSSION Considerable efforts have been made in the past by several research groups to study the mechanisms involved in functionalized surfactant-assisted hydrolysis of phosphate esters.48-53 Buncel et al. have studied the effect of a series of oximates on the degradation of pesticide fenitrothion in cationic micellar media.54Several amphiphilic quaternary pyridinium ketoximes have been synthesized by Hampl et al.55and evaluated for their ability to cleave phosphate esters. Simanenko et al. developed new class of functionalized surfactants with imidazolium and pyridinium rings. Both the rings were linked to a variety of α-nucleophilic moieties varying from oxime, hydroxamic to amidoxime. The group further studied their detailed kinetic analysis for the decomposition of phosphate ester.56-59 The enhancement in the rate of degradation of phosphoesters by oximate ions in presence of cationic surfactants was well documented by Bunton60, Toullec61 and co-workers. Further, Bunton et al., have reported quaternized aliphatic aldxoimes to study the

11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 39

size versus reactivity of functionalized surfactants towards the dephosphorylation and deacylation of phosphate esters.62 Bhattacharya and research group discussed the synthesis and roles of metallosurfactants in the hydrolysis of phosphate, carboxylate and sulphonate esters.63 Recently, Renard et al. reported a series of oximes and amidoximes based α-nucleophiles and also investigated their ability to cleave P=S bond of OP-based chemical warfare agents.64 It is a wellknown fact that micelles and other self-organized assemblies boost the observed rates of nucleophilic processes.65 In the present investigation, we have contemplated the physicochemical and surface properties of conventional and oxime based functionalized surfactants; F.S. (Figure 2) in both aqueous and buffer media. The strength of F.S. as both a nucleophilic and catalytic self-organized assembly for efficient esterolytic reactions has been explored. Physicochemical properties of cationic functionalized 3- and 4-hydroxyiminomethyl-1-alkylpyridinium bromide surfactants (3Cn and 4-Cn) were investigated in both the presence and absence of conventional cationic surfactants cetylpyridinium bromide (CPB), cetyltrimethylammonium bromide (CTAB) and cetyldimethylethanolammonium bromide (CDMEAB). The mixed micellar systems of oximefunctionalized and conventional surfactants were also used to study the hydrolysis of tested phosphate esters. The pseudo first order rate constants were determined for each kinetic run under identical conditions. Micellar and Surface Properties The critical micelle concentration; CMC and other parameters of pure and mixed (conventional surfactants + 3-Cn (n = 10, 12, 14, 16) and 4-Cn (n = 10, 12) systems were determined by surface tension measurements (presented in Table 1) at pH 7.0. CMC of all the investigated functionalized surfactants of 3-Cn series in the mixed system of CPB are presented in Figure 5. It can be clearly

12 ACS Paragon Plus Environment

Page 13 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

observed from Table 1 that the CMC of pure components and mixed systems depend on hydrophobic chain length of the surfactants. With increase in alkyl chain length, CMC values of the surfactants decrease significantly in the order 3-C10 > 3-C12 > 3-C14 > 3-C16. Hydrophobicity of surfactants and polarity of bulk phase usually determined the extent of micellization. Increment in the alkyl chain lengths enhances the hydrophobicity of surfactants results into optimum thermodynamic threshold required for the formation of micelle. This threshold value is lower in case of surfactant with higher chain. CMC value of mixed system is lower in comparison to that of the individual components.66 The ideality of mixed micelles system which represents the correlation of theoretical ideal CMC value with the experimental CMC value of pure components is determined by using following Clint equation:

Where CMCideal and CMCi are the critical micelle concentration of the mixture and ith component, respectively; αi is the mole fraction of component i in solution. From Table 1 it is concluded that CMCi value is higher than CMCm value throughout the alkyl chain except in case of 3-C16+ CTAB and 3-C16 + CDMEAB. This suggests that the micellization behavior is throughout synergistic except 3-C16+ CTAB and 3-C16 + CDMEAB which showed antagonistic behavior.

13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 39

Figure 5. Plots of the surface tension vs log C of conventional surfactants + functionalized surfactants. Acid Dissociation Constant (pKa) As the oximate anion involves in the hydrolysis reaction, it was necessary to determine the pKa of the oximes. In our previous works we have reported the pKa of the 3- and 4-series of hydroxyiminomethyl-1-alkylpyridinium bromide functionalized surfactants and their role in the hydrolytic transformations toward dephosphorylation of di- and triphosphate esters.41, 43, 67 The effect of surfactant concentration on the pKa of nucleophiles is essential for analyzing the quantitative effects of surfactant assemblies on the rate of esterolysis. The calculated pKa values of the functionalized surfactants are in well agreement with the literature value. To assess pKa under defined reaction conditions, we studied the effect of surfactants (CPB, CTAB, CDMEAB) concentration on the acid dissociation constant value of the

14 ACS Paragon Plus Environment

Page 15 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Sustainable Chemistry & Engineering

Table 1. The critical micelle concentration (CMC), maximum surface excess concentration at the air/water interface (Гmax), minimum area per surfactant molecule at the air-water interface (Amin) and the surface pressure at the CMC (πCMC) by surface tension; pH 7.00, 27˚C. F.S.

CMC (mM) F.S.

F.S. + CPB (1:1) CMCm

CMCi

(mM)

max (106

F.S. + CTAB (1:1) ∏CMC

Amin 18

2

(10 m )

-1

(mNm )

CMCm

CMCi

(mM)

max (106

mol m-2)

F.S. + CDMEAB (1:1) ∏CMC

Amin 18

2

(10 m )

-1

(mN m )

CMCm

CMCi

(mM)

mol m-2)

max

Amin

∏CMC 18

(106

(10

mol m-2)

m2)

(mN m-1)

3-C10

2.03a

0.55

1.32

1.287

1.290

25.5

0.80

1.24

1.239

1.340

26

0.70

1.15

1.308

1.269

27

3-C12

1.20a

0.45

1.08

1.914

0.867

26.0

0.70

1.03

1.329

1.248

23

0.65

0.96

0.949

1.749

25

3-C14

0.65a

0.35

0.78

2.057

0.807

27.0

0.75

0.75

1.264

1.313

24

0.70

0.72

1.058

1.569

21

3-C16

0.08a

0.09

0.15

1.866

0.889

29.0

0.50

0.14

2.119

0.783

30

0.45

0.14

2.262

0.734

30

a=55

ACS Paragon Plus Environment

15

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 39

investigated functionalized surfactants by spectrophotometric techniques. Scheme 2 was applied for pyridinium-based oxime functionalized surfactants:

Scheme 2. Deprotonation of pyridinium oxime It can be concluded from Table 2 that (i) the pKa value of functionalized surfactants decrease with an increase in alkyl chain length (ii) pKa value of 4-series functionalized surfactants are lower in comparison to 3-series and can be well accounted on the basis of spatial hindrance offered by the oxime moiety at 3- position in the pyridinium ring, and (iii) pKa values decrease in the presence of used monomeric surfactants (CPB, CTAB and CDMEAB). To support this fact, pKa of 3-C16 was studied at various CPB concentrations (Figure 6). At higher CPB concentration, considerable decrease in pKa value was observed. This can be well explained by the fact that, at concentration lower than the CMC value of conventional surfactants, decrease in pKa value is not much significant,68 but at high concentration of surfactants the total CMC of the binary system is reduced resulted in shifting of acid-base equilibrium. Charged micelles enhance the redistribution of acidic and basic species into the micellar pseudophase. pKa values of functionalized surfactant in the presence and in absence of conventional surfactants strongly support the shifts in acid-base equilibrium.69 The shifts of apparent constants (pKa, app) values in the presence of micelles have been well discussed by Berezin and Yatsimirsy in the framework of pseudophase partitioning model (PPM) of micellar catalysis,70 and well developed by Romsted, Bunton and others in the framework of pseudophase ion exchange (PIE) approach.65,71

16 ACS Paragon Plus Environment

Page 17 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Sustainable Chemistry & Engineering

Table 2. Acid dissociation constants (pKa) values of 3-Cn series of functionalized surfactants in the presence and absence of conventional surfactants viz. CPB, CTAB and CDMEAB by Albert and Sergeant method47; 27˚C 3-C10 3-C12 3-C14 3-C16 Conc. CPB CTAB CDMEAB CPB CTAB CDMEAB CPB CTAB CDMEAB CPB CTAB CDMEAB (mM) 0.0

9.40

9.40

9.40

9.32

9.32

9.32

9.22

9.22

9.22

8.68

8.68

8.68

0.50

9.20

9.33

9.36

9.21

9.26

9.29

9.05

9.05

9.12

8.56

8.62

8.65

1.00

9.02

9.23

9.27

8.80

9.03

9.07

8.98

9.08

9.14

8.50

8.58

8.60

2.00

8.82

9.18

9.23

8.75

8.90

8.95

8.62

8.98

9.06

8.40

8.51

8.54

3.00

8.72

9.02

9.11

8.35

8.74

8.80

8.51

8.86

8.96

8.28

8.43

8.48

4.00

8.60

8.85

9.01

8.23

8.45

8.64

8.46

8.76

8.88

8.15

8.36

8.42

ACS Paragon Plus Environment

17

ACS Sustainable Chemistry & Engineering

a 8,8

CPB CTAB CDMEAB

pKa

8,6

8,4

8,2

8 0

1

2

3

4

5

Inert surfactant /mM

b 8,8

CPB CTAB 8,6

pKa

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 39

CDMEAB

8,4

8,2

8 0

1

2

3

4

5

Inert surfactant /mM

Figure 6. Acid dissociation constant (pKa) of functionalized surfactants at various concentrations of inert surfactants. a - 3-hydroxyiminomethyl-1-hexadecylpyridinium bromide (3-C16); b - 4hydroxyiminomethyl-1-dodecylpyridinium bromide (4-C12).

18 ACS Paragon Plus Environment

Page 19 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Within the PIE model it was assumed that the positively charged micelle binds OH- and both protonated and deprotonated forms of a species, increasing deprotonation and decreasing acidity constant. At the same time, surfactant counter ion tends to competitively displace negatively charged form of a nucleophile from the micellar surface. This model leads to a very simple quantitative treatment for both normal O-nucleophiles and oximes. The apparent basicity constants estimated in the micellar pseudophase was shown to be larger than those in water up to order of magnitude.71 Although some other approaches have been suggested72 for the estimation of differentiating effect of the colloid systems on the acid-base equilibria, the PIE remains most useful tool for the estimations of pKa app shifts in the presence of ionic micelles. Study of pH effect allowed us to determine the kinetic pKa value of nucleophiles. The effect of pH was monitored on hydrolysis of paraoxon with 0.5 mM of 4-C12 in the presence and absence of CPB (1:1) by taking different buffers (0.1 M) ranging from pH 6 to 10 as added to the reaction medium. It was observed that the rate of reaction is increased with the increase in pH value. A pH versus rate constant profile for the nucleophilic cleavage of NPDEP by 4-C12 oximate anion gave the apparent pKa value (8.45) for the 4-C12 pyridinium aldoxime as represented in Figure 7 a. The log kobs vs pH plots for 4-C12 revealed that oximate ions from the functional surfactants along with the co-micellar environment of CPB are more reactive and triggered considerable reduction in pKa (Figure 7). 3-C16 + CPB proved to be the most efficient system for degradation of phosphate esters and this has been demonstrated by the strong synergistic interactions obtained by surface tension measurements, see Table 1.

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 39

Figure 7. Profile log kobs vs pH for the cleavage of NPDEP with 4-C12 in the presence and absence of CPB. Inset: pH profiles for the cleavage of NPDEP with 4-C12. Reaction conditions: [4-C12] = 0.5 mM, [NPDEP] = 0.5 mM, μ = 0.1 M (KCl), [CPB] = 1.0 mM; 27˚C. Micellar hydrolysis with functionalized surfactants Among the various pyridinium and imidazolium based functionalized surfactants, pyridinium functionalized surfactants proved to be an efficient nucleophile for the effective cleavage of various phosphate and sulphonate esters in co-micellar medium. We have performed the hydrolysis of both the pesticides using 3-Cn and 4-Cn series of pyridinium aldoxime based nucleophiles in the co-micellar medium of the conventional surfactants. Studies could not be performed on the 4substituted pyridinium functionalized surfactants having C14 and C16 alkyl chain lengths due the solubility problems in the reaction mixture. Effect of alkyl chain length of functionalized surfactants Table 3 and Table 4 show the effect of alkyl chain lengths (C10, C12, C14, and C16) of the 3-Cn functionalized surfactants on the esterolytic reactions. Alkyl chain length has a significant impact

20 ACS Paragon Plus Environment

Page 21 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

on the kobs values for the cleavage of phosphate esters of pesticides. The kobs values increased with increase in the alkyl chain length of the surfactants. This increase was in the order of C10 < C12 < C14 < C16. The observed trend may be accounted on the basis of enhancement in hydrophobicity of the palisade layer of micelle and to the increment in electrical surface potential of the micelles in the stated order. The rate increase is strongly supported by the observed CMC and pKa of the studied F.S. The pKa value of 3-C16 (8.50) is the least (3-C10, 9.02; 3-C12, 8.80; and 3-C14, 9.08) therefore, 3-C16 showed maximum reactivity as nucleophilic moiety in micellar media, see Figure 8. Also, CMCs of these functionalized surfactants are in the order 3-C10 > 3-C12 > 3-C14 > 3-C16, which clearly suggests higher reaction rates with the rest of the 3-Cn nucleophiles, similarly reactivity order is in case of 4-Cn is 4-C12>4-C10 (Table S1).

(a)

(b)

Figure 8. Rate profile for effect of CPB as comicelle on the hydrolysis of (a) NPDEP and (b) NPDMP with 3-Cn (n= 10, 12, 14, 16). Reaction conditions: [Substrate] = 0.05 mM, [Nu-] = 0.5 mM, μ = 0.1 M (KCl); 27˚C. The observed trend of alkyl chain lengths of 3-Cn in Table 5 and Table 6 is also supported by the results obtained for the effect of conventional surfactants as mixed micelles on rate of phosphate esters cleavage.

21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 22 of 39

Table 3. Pseudo-first-order rate constants for the hydrolysis of NPDEP with 3-Cn F.S. Conc. (mM)

CPB

CTAB

CDMEAB

3C10

3C12

3C14

3C16

3C10

3C12

3C14

3C16

3C10

3C12

3C14

3C16

0

0.002

0.008

0.275

0.587

0.002

0.008

0.275

0.587

0.002 0.008

0.275

0.587

0.25

0.069

0.173

0.990

1.320

0.095

0.200

0.513

1.120

0.088 0.126

0.576

0.889

0.50

0.191

0.250

1.300

1.860

0.198

0.266

1.030

1.650

0.138 0.168

0.892

1.550

1.0

0.447

0.851

2.010

2.640

0.238

0.498

1.850

2.520

0.215 0.248

1.590

2.450

1.5

0.512

0.901

2.260

3.200

0.198

0.540

2.100

2.850

0.284 0.318

2.120

2.950

2.0

0.417

0.782

2.160

3.040

0.184

0.486

2.080

2.780

0.312 0.380

1.970

2.720

2.5

0.346

0.731

1.980

2.860

0.153

0.455

1.910

2.550

0.305 0.401

1.520

2.210

3.0

0.253

0.620

1.890

2.420

0.121

0.410

1.830

2.380

0.298 0.393

1.210

1.840

3.5

0.219

0.4711

1.510

2.100

0.11

0.381

1.640

2.170

0.275 0.376

1.010

1.690

4.0

0.200

0.465

1.310

1.970

0.104

0.312

1.540

2.020

0.278 0.362

0.913

1.490

5.0

0.190

0.456

1.290

1.910

0.101

0.296

1.430

1.930

0.281 0.371

0.891

1.310

Reaction conditions: [Nu-] = 0.5 mM, [substrate] = 0.05 mM,  = 0.1 M (KCl), pH = 9.0; 27˚C.

ACS Paragon Plus Environment

22

Page 23 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

ACS Sustainable Chemistry & Engineering

Table 4. Pseudo-first-order rate constants for the hydrolysis of NPDMP with F.S. Conc. (mM)

3-C10

CPB 3-C12 3-C14

3-C10

CTAB 3-C12 3-C14

3-C16

0

0.002

0.008

0.126

0.25

0.098

0.165

0.50

0.156

1.0

CDMEAB 3-C12 3-C14

3-C16

3-C10

0.206

0.002

0.008

0.126

0.206

0.002 0.008

0.126

0.206

0.512

0.786

0.073

0.139

0.458

0.662

0.088 0.126

0.326

0.583

0.376

0.986

1.550

0.195

0.315

0.714

1.180

0.138 0.168

0.621

0.984

0.434

0.649

1.710

2.280

0.352

0.526

1.190

1.670

0.284 0.248

0..996

1.360

1.5

0.486

0.691

2.030

2.710

0.391

0.609

1.440

1.920

0.322 0.401

1.290

1.710

2.0

0.384

0.568

1.840

2.560

0.355

0.598

1.390

1.810

0.305 0.318

1.190

1.590

2.5

0.295

0.481

1.620

2.320

0.309

0.506

1.290

1.720

0.298 0.380

1.010

1.380

3.0

0.275

0.429

1.490

1.920

0.231

0.418

1.190

1.540

0.281 0.393

0.950

1.120

3.5

0.248

0.401

1.400

1.730

0.186

0.392

1.090

1.490

0.275 0.376

0.900

0.996

4.0

0.210

0.365

1.330

1.650

0.144

0.333

1.020

1.380

0.278 0.362

0.891

0.982

5.0

0.195

0.326

1.290

1.600

0.121

0.303

0..998

1.310

0.271 0.371

0.874

0.961

3-C16

Reaction conditions: [Nu-] = 0.5 mM, [substrate] = 0.05 mM,  = 0.1 M (KCl), pH = 9.0; 27˚C.

ACS Paragon Plus Environment

23

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 39

Effect of Head Group The effect of hydrophilic head group of conventional cationic surfactants bearing same cetyl chain length was also investigated on the hydrolysis of NPDEP with 4-C12 and is presented in Figure 9.

Figure 9. Rate-surfactant concentration profile for head group effect of conventional cationic surfactants for the hydrolysis of NPDEP with 3-C16. Reaction conditions: [Substrate] = 0.05 mM, [Nu-] = 0.5 mM, μ = 0.1 M (KCl); pH = 9.0; 27˚C. The comparison of reaction rate was made between the surfactant bearing head groups varying by aromatic (CPB) and aliphatic (CTAB and CDMEAB). It can be observed under comparable conditions, that the rate constants (kobs) for the hydrolysis of phosphate esters were found to be maximum with surfactants bearing aromatic ring as head group than for aliphatic nitrogen centers. Reactivity order for micellar hydrolysis of cationic surfactants was obtained as CPB > CTAB > CDMEAB. Presence of bulky head groups may be the probable reason for comparative rate enhancement in the presence of pyridinium based cationic surfactants than alkyl ammonium head groups. The effect led to a decrease in the polarity of the head group region which therefore increased the rate of the nucleophilic substitution. The enhancement in interruption of

24 ACS Paragon Plus Environment

Page 25 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

the bromide ion on the bulky head group results in the decrease of the polarity of the head group region consequently increment in the nucleophilic substitution observed. The half-lives of paraoxon in water and different nucleophilic systems, including “normal” nucleophilic reagents e.g. H2O (spontaneous, or neutral hydrolysis), hydroxide ion (alkaline hydrolysis), several alpha effect nucleophiles e.g. hypobromite, iodozobenzoate ions, and oxime functionalized surfactants (pyridinium aldoximes reported in this paper with pyridinium ketoximes and imidazolium oximes published earlier) are collected in the Table 5. Table 5 Half-lives of pesticide paraoxon (NPDEP) in different nucleophilic reaction systems. Nucleophile



Experimental conditions

H 2O

≥50 d

pH 7.073

OH-

>7 d 20 h 6d 19 h

pH 9.3 pH 10.574

BrO- a

1750 s 75, 76 770 s 76 200 s 25 380 s 230 s 900 s 420 s 1100 s 610 s 150 s 25 300 s

215 s

pH 9.3; [CTAB] = 3mM pH 10.5; [CTAB] = 3mM pH 11.35; [HOBr]0 = 40 mM; [KCl] = 1.0 M pH 11.15; [CTABr3] = 20 mM R = C16H33; pH >10.0; [FS+CTABr] >>20 mM R = C10H21; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CPB] = 2 mM R = C12H25; [FS] = 0.5 mM, [KCl] = 0.1 M, pH = 9.0; [CPB] = 2 mM R = C10H21; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CTAB] = 2 mM R = C12H25; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CTAB] = 2 mM R = C10H21; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CDMEAB] = 2 mM R = C12H25; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CDMEAB] = 2 mM R = C16H33; pH>10.6; [FS+CTABr] >> 20 mM R = C14H29; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CPB] = 2 mM R= C16H33; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CPB] = 2 mM

25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

330 s 245 s 325 s 77 235 s 77

Page 26 of 39

R = C14H29; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CTAB] = 2 mM R= C16H33; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CTAB] = 2 mM R = C14H29; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CDMEAB] = 2 mM R= C16H33; [FS] = 0.5 mM, [KCl] = 0.1 M, pH 9.0; [CDMEAB] = 2 mM

190 s 78

pH 10.5; R= C12H25; [FS] = 12 mM

410 s 78

pH 10.5; R= C14H29; [FS] = 1.5 mM; [CTAB] = 4.5 mM pH 10.5; R= C16H33; [FS] = 0.5 mM, [CTAB] = 4.5 mM

295 s 78 180 s77

pH 8.0; [CTACl] = [Nu]0 = 20 mM; [Tris] = 0.1 M

200 s77

pH 8.0; [CTACl] = 20 mM; [FS]= 4 mM [KCl] = 0.1 M; [Tris] = 0.1 M

46 s 79

R = C12H25; [FS] = 10 mM, pH = 10.50.

28 s 79

R = C14H29; [FS] = 10 mM, pH = 10.50.

60 s58 92 s29

pH 12.90; [FS] = 60 mM pH 11.42; [FS] = 42 mM

50 s29

pH 10.94; [FS] = 13.4 mM

Notes. a BrO- generated from Br2 or stable active bromine sources reported in75-76. Functionalized surfactants designed as FS

The Scheme 1 presents the pivotal stage of OP detoxification, namely nucleophilic attack of an oximate ion at the electrophilic phosphorus center. Further transformation includes hydrolysis of phosphorylated oxime, one of the products to be acid of the parent ester (OP substrate) having incomparably lower toxicity than the initial ester. In contrast to inorganic alpha

26 ACS Paragon Plus Environment

Page 27 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

nucleophiles, hypohalite or hydroperoxide ions acting as “true” catalysts,80,81 oximates demonstrate more complex behavior in the transformation of acylated (prosphorylated) intermediate. As reported elsewhere,82,83,84 the second product of an O-acyloxime hydrolysis can be initial oxime (catalytic route), or no oxime generation occurs (stoichiometric route). The two routes can coexist with variable yields, and in the case of non-catalytic route, N-heterocyclic oximes generate nitriles (except the 2-pyridinium compounds,85 irrespectively of the electrophilic center (phosphate, phosphonate, alkylcaboxylate). Despite developing catalytic oxime-containing functional micellar system is among the most challenging issues in designing decontamination formulations,65, 86 some remarkable examples of turnover systems are well-known.61, 87 Yield of the catalytic pathway depends on several factors including basicity of an aldoxime (or a ketoxime) and properties of electrophilic center. The environmental fate of organic compounds is dependent on several factors, e.g. whether this particular compound and its transformation products are persistent and whether they are toxic to the environment 88. The latter factor prevents it to be readily biodegradable even if it is not toxic, and the latter does not allow the microorganisms to transform it or its metabolites to the nontoxic (preferably mineralizable) products. In spite of the fact that, real time of environmental fate of organophosphorus pesticides is dependent on the type of ground water (pH, salinity, temperature, etc.)89 and therefore can be different from the spontaneous hydrolysis found in the laboratory,73 see Table 5, half lives are big enough to have a negative impact on the environment. It is the toxicity of paraoxon which prevents its degradation and thus hydrolytic cleavage 90-91 remains the main method to overcome the “toxicity barrier”. Current interest of the researchers and society to the pesticides degradation at low concentrations and in low-nutrient situations (groundwater, in essence) requires an efficient array of abiotic and biotic transformations effectively removing

27 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 39

pesticides from the environment.92 The pyridinium-based surfactants can provide reducing halflife of insecticide paraoxon to minutes (see Table 5) in water under mild pH (below pH 10) and concentration condition. As was recently shown for structurally similar 2-oximinopyridinium surfactants93, the microbial toxicity of the functional surfactants is not dramatically different from those having pyridinium headgroup without nucleophilic moiety. The “benign by design” approach developed recently by Kummerer and Gathergood 38, 39 may open an opportunity for using degradable functionalized pyridinium surfactants and ionic liquids as the base for sustainable decontamination formulations. Conclusions In present investigation, an attempt has been made to investigate the kinetic efficiency of pyridinium based functionalized surfactants towards the micellar hydrolysis of organophosphate based pesticides. Physicochemical and surface properties of individual and mixed (binary) system were thoroughly studied. Considerable enhancement of rate of hydrolysis reaction was observed on increment of the alkyl chain length of functionalized surfactants. 3C16 showed maximum reactivity because of lower acid dissociation constant (pKa) and CMC value. Mixed micellar systems composed by functionalized and conventional surfactants accelerate pesticides hydrolysis. The first-order rate constant increased up to a certain concentration then remained constant or dropped down. 3-C16 / CPB proved to be the most efficient system for degradation of phosphate esters as this system showed highest synergistic interactions. Results of the present investigation may provide useful information for development of an effective system for degradation of toxic pesticides and nerve agents. Since a unique and broad-spectrum cleaving agent is still lacking, the obtained results can be utilized in the development of the sustainable formulations for degradation of organophosphorus pesticides.

28 ACS Paragon Plus Environment

Page 29 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Notes The authors declare no competing financial interest. Acknowledgments Financial assistance from DRDO (New Delhi), project (No.ERIP/ER/1003906/M/01/1393) is gratefully acknowledged. Authors are also thankful to DST-FIST [No.SR/FST/CSI-259/2014(C)] for financial support. NG and YK acknowledge funding from the European Union’s Seventh Framework Programme for research, technological development and demonstration activities under grant agreement No. 621364 (TUTIC-Green). YK and KK thank excellence project of Faculty of Informatics and Management, University of Hradec Kralove (UHHK). KK and JM thank to Long-term development plan UHHK. JM thanks to long-term development project of FVZ UO and FM OU. Supporting Information Acid dissociation constants of 4-Cn series of functionalized surfactants (F.S.) in the presence and absence of the conventional surfactants in aqueous medium at 300K; surface properties of 3-Cn series of F.S. in absence and presence of conventional surfactants at pH 8.00 and pH 9.00; effect of conventional surfactants as comicelle on the hydrolysis of NPDEP and NPDMP with 3-Cn and 4-Cn F.S.; rate profile for the esterolytic reaction of NPDEP and NPDMP with 3-Cn and 4-Cn in the presence of conventional surfactants. This material is available free of charge via the Internet at http://pubs.acs.org.

Funding Sources

29 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 39

Financial assistance from DRDO (New Delhi), project (No.ERIP/ER/1003906/M/01/1393), DSTFIST [No. SR/FST/CSI-259/2014(C)], and FP7 [award No. 621364, TUTIC-Green] are greatly acknowledged. Author Contributions. The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

REFERENCES (1)

Environmental Protection Strategies for Sustainable Development, Malik, A.; Grohmann, E. (eds.), Springer Science Business Media B.V. 2012, 605.

(2)

Kirby, A. J.; Warren, S. G. The Organic Chemistry of Phosphorus; Elsevier: Amsterdam, 1967.

(3)

Eddleston, M.; Buckley, N. A.; Eyer, P.; Dawson, A. H. Management of acute organophosphorus pesticide poisoning. Lancet 2008, 371 (9612), 597-607.

(4)

Costa, L.G. Current issues in organophosphate toxicology. Clin. Chim. Acta. 2006, 366 (1-2), 1-13.

(5)

Mercey G.; Verdelet, T.; Renou, J.; Kliachyna, M.; Baati, R.; Nachon, F.; Jean L, Renard, P.Y. Reactivators of acetylcholinesterase inhibited by organophosphorus nerve agents. Acc. Chem. Res., 2012, 45 (5), 756-766.

(6)

Sharma, R.; Gupta, B.; Singh, N.; Acharya, J. R.; Musilek, K.; Kuca, K.; Ghosh, K. K. Development and Structural Modifications of Cholinesterase Reactivators against Chemical Warfare Agents in Last Decade: A Review. Mini. Rev. Med. Chem. 2015, 15 (1), 58-72.

(7)

Okumura, T.; Hisaoka, T.; Yamada, A.; Naito, T.; Isonuma, H.; Okumura, S.; Miura, K.; Sakurada, M.; Maekawa, H.; Ishimatsu, S.; Takasu, N.; Suzuki, K. The Tokyo subway sarin attack—lessons learned. Toxicol. Appl. Pharmacol. 2005, 207 (2), 471–476.

(8)

Rosman, Y.; Eisenkraft, A.; Milk, N.; Shiyovich, A.; Ophir, N.; Shrot, S.; Kreiss, Y.; Kassirer, M. Lessons learned from the Syrian sarin attack: evaluation of a clinical syndrome through social media, Ann. Intern. Med. 2014, 160 (9), 644–648.

30 ACS Paragon Plus Environment

Page 31 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(9)

Gould, Ch., Folb, P.I., Project Coast: Apartheid’s Chemical and Biological Warfare Programme; Berold R, Ed. United Nations Publications UNIDIR – CRR: Geneva – Cape Town, 2002.

(10)

Hollingworth, R.M. The Biochemical and Physiological Basis of Selective Toxicity. In Insecticide Biochemistry and Physiology; Wilkinson, C. F., Eds.; Springer Science + Business Media, New York, NY, U.S.A, 1976.

(11)

Kardos, S. A.; Sultatos, L. G. Interactions of the Organophosphates Paraoxon and Methyl Paraoxon with Mouse Brain Acetylcholinesterase. Toxicol Sci. 2000, 58 (1), 118-126.

(12)

Petroianu, G. A.; Lorke, D. E.; Kalasz, H. Comparison of the Ability of Pyridinium Aldoximes to Reactivate Human Red Blood Cell Acetylcholinesterases Inhibited by Ethyl- and Methyl-Paraoxon. Curr. Org. Chem., 2012, 16 (10), 1359-1369.

(13)

Musilek, K.; Komloova, M.; Holas, O.; Horova, A.; Pohanka, M.; Moore, F. G.; Dohnal, V.; Dolezal, M.; Kuca, K. Mono-Oxime Bisquaternary Acetylcholinesterase Reactivators with Prop-1,3-diyl Linkage-Preparation, In-Vitro Screening and Molecular Docking. Bioorg. Med. Chem. 2011, 19 (2), 754−762.

(14)

Singh. N.; Karpichev. Y.; Tiwari, A. K.; Kuca, K.; Ghosh, K. K. Oxime Functionality in Surfactant Self-assembly: An Overview on Combating Toxicity of Organophosphates. J. Mol. Liq. 2015, 208, 237-252.

(15)

Satnami, M. L.; Dhritlahre, S.; Nagwanshi, R.; Karbhal, I.; Ghosh, K. K.; Nome, F. Nucleophilic Attack of Salicylhydroxamate Ion at C=O and P=O Centers in Cationic Micellar Media. J. Phys. Chem. B. 2010, 114 (50), 16759−16765.

(16)

Ghosh, K. K.; Sinha, D.; Satnami, M. L.; Dubey, D. K.; Dafonte, P. R.; Mundhara, G. L. Nucleophilic Dephosphorylation of p-Nitrophenyl Diphenyl Phosphate in Cationic Micellar Media. Langmuir 2005, 21 (19), 8664−8669.

(17)

Orth, E. S.; Silva, P. L. F.; Mello, R. S.; Bunton, C. A.; Milagre, H. M. S.; Eberlin, M. N.; Fiedler, H. D.; Nome, F. Suicide Nucleophilic Attack: Reactions of Benzohydroxamate Anion with Bis(2,4-dinitrophenyl) Phosphate. J. Org. Chem. 2009, 74 (14), 5011−5016.

(18)

Singh, N.; Karpichev, Y.; Sharma, R.; Gupta, B.; Sahu, A. K.; Satnami, M. L.; Ghosh, K. K. From Alpha Nucleophiles to Functionalized Aggregates: Exploring the Reactivity of

31 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 39

Hydroxamate Ion towards Esterolytic Reactions in Micelles. Org. Biomol. Chem. 2015, 13 (10), 2827-2848. (19)

Kirby, A. J.; Manfredi, A. M.; Souza, B. S.; Medeiros, M.; Priebe, J. P.; Brandao, T. A. S.; Nome, F. Reactions of Alpha Nucleophiles with a Model Phosphate Diester. ARKIVOC , 2009 (3), 28−38.

(20)

Morales-Rojas, H. M.; Moss, R. A. Phosphorolytic Reactivity of O-iodosylcarboxylates and Related Nucleophiles. Chem. Rev. 2002, 102 (7), 2497−2521.

(21)

Jokanović, M. Pyridinium Oximes in the Treatment of Poisoning with Organophosphorus Compounds, Ch. 71. In Handbook of Toxicology of Chemical Warfare Agents (2nd Ed.), R. C. Gupta, Ed. - Academic Press. - 2015, 1057–1070.

(22)

Sirieix, J.; Viguerie, N.; Rivière, M.; Lattes, A. Amphiphilic Urocanic Acid Derivatives as Catalysts of Ester Hydrolysis. New J. Chem. 1999, 23 (1), 103−109.

(23)

Liu, X.; Dong, L.; Fang, Y. Synthesis and Self-aggregation of a Hydroxyl-Functionalized Imidazolium-Based Ionic Liquid Surfactant in Aqueous Solution. J. Surfactants Deterg. 2011, 14 (2), 203−210.

(24)

Tonellato, U. Reactivity in Functionalized Assemblies. Colloids Surf. 1989, 35 (2), 121−134.

(25)

Kapitanov, I. V.; Belousova, I. A.; Turovskaya, M. K.; Karpichev, E. A.; Prokop’eva, T. M.; Popov, A. F. Reactivity of Micellar Systems Based on Supernucleophilic Functional Surfactants in Processes of Acyl Group Transfer. Russ. J. Org. Chem. 2012, 48 (5), 651−662.

(26)

Kotoucová, H.; Cibulka, R.; Hampl, F.; Liška, F. Amphiphilic Quaternary Pyridinium Ketoximes as Functional Hydrolytic Micellar Catalysts-Does the Nucleophilic Function Position Influence Their Reactivity? J. Mol. Catal. A: Chem. 2001, 174 (1-2), 59−62.

(27)

Kivala, M.; Cibulka, R.; Hampl, F. Cleavage of 4-Nitrophenyl Diphenyl Phosphate by Quaternary Pyridinium Ketoximes-How Can Structure and Lipophilicity of Functional Surfactants Influence Their Reactivity in Micelles and Microemulsions? Collect. Czech. Chem. Commun. 2006, 71 (11-12), 1642−1658.

(28)

Scrimin, P.; Tecilla, P.; Tonellato, U.; Bunton, C. A. Nucleophilic Catalysis of Hydrolysis of Phosphate and Carboxylate Esters by Metallomicelles. Facts and Misconceptions. Colloids Surf. A 1998, 144 (1-3), 71−79.

32 ACS Paragon Plus Environment

Page 33 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(29)

Kapitanov, I. V.; Belousova, I. A.; Shumeiko, A. E.; Kostrikin, M. L.; Prokop’eva, T. M.; Popov, A. F. Supernucleophilic systems based on functionalized surfactants in the decomposition of 4-nitrophenyl esters derived from phosphorus and sulfur acids: I. Reactivity of a hydroxyimino derivative of gemini imidazolium surfactant. Russ. J. Org. Chem. 2013, 49 (9), 1291-1299.

(30)

Plechkova, N.V.; Seddon, K.R. Applications of ionic liquids in the chemical industry. Chem. Soc. Rev. 2008, 37(1),123−150.

(31)

Tian, F.; Liu, W.; Fang, H.; An, M.; Duan, S. Determination of Six Organophosphorus Pesticides in Water by Single-Drop Microextraction Coupled with GC-NPD. Chromatographia 2014, 77 (5), 487–492.

(32)

Zheng, Y.; Liu, Zh.; Zhan, H.; Li, J.; Zhang, Ch. Development of a Sensitive Acetylcholinesterase Biosensor Based on a Functionalized Graphene–Polyvinyl Alcohol Nanocomposite for Organophosphorous Pesticide Detection. Anal. Methods, 2015, 7 (23), 9977-9983.

(33)

Fan, T.; Wu, X.; Peng, Q. Sparingly Soluble Pesticide Dissolved in Ionic Liquid Aqueous. J. Phys. Chem. B 2014, 118 (39), 11546−11551.

(34)

Fan, T.; Chen, Ch.; Fan, T.; Liu, F.; Peng, Q. Novel Surface-Active Ionic Liquids Used as Solubilizers for Water-Insoluble Pesticides. J. Hazard. Mater. 2015, 297, 340–346.

(35)

Bica, K.; Gartner, P.; Gritsch, Ph. J.; Ressmann, A. K.; Schroder, Ch.; Zirbs, R. Micellar Catalysis in Aqueous–Ionic Liquid Systems. Chem. Commun., 2012, 48 (41), 5013–5015.

(36)

Jordan, A.; Gathergood, N. Biodegradation of ionic liquids – a critical review. Chem. Soc. Rev., 2015, 44 (22), 8200-8237.

(37)

Sharma, R.; Gupta, B.; Sahu, A. K.; Musilek, K.; Kuca, K.; Acharya, J. R.; Ghosh, K. K. Synthesis and in-vitro reactivation screening of imidazolium aldoximes as reactivators of sarin and VX-inhibited human acetylcholinesterase (hAChE). Chem. Biol. Interact., 2016, doi:10.1016/j.cbi.2016.04.034.

(38)

Haiß, A.; Jordan, A.; Westphal, J.; Logunova, E.; Gathergood, N.; Kümmerer, K. On The Way to Greener Ionic Liquids: Identification of a Fully Mineralizable PhenylalanineBased Ionic Liquid. Green Chem., 2016, 18, 4361-4373.

(39)

Jordan, A.; Haiß, A.; Spulak, M.; Karpichev, Y.; Kümmerer, K.; Gathergood, N. Synthesis of a Series of Amino Acid Derived Ionic Liquids and Tertiary Amines: Green

33 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 39

Chemistry Metrics Including Microbial Toxicity and Preliminary Biodegradation Data Analysis. Green Chem., 2016, 18, 4374-4392. (40)

Ghosh, K. K.; Sinha, D.; Satnami, M. L.; Dubey, D. K.; Shrivastava, A.; Palepu, R. M.; Dafonte, P. R. Enhanced Nucleophilic Reactivity of Hydroxamate Ions In Some Novel Micellar Systems For The Cleavage Of Parathion. J. Colloid Interface Sci. 2006, 301 (2), 564-568.

(41)

Singh, N.; Karpichev, Y.; Gupta, B.; Satnami, M. L.; Marek, J.; Kuca, K.; Ghosh, K. K. Physicochemical Properties and Supernucleophilicity of Oxime-Functionalized Surfactants: Hydrolytic Catalysts toward Dephosphorylation of Di- and Triphosphate Esters. J. Phys. Chem. B 2013, 117 (14), 3806−3817.

(42)

Singh, N.; Ghosh, K.K; Marek, J.; Kuca, K. Hydrolysis of carboxylate and phosphate esters using monopyridinium oximes in cationic micellar media. Int. J. Chem. Kinet. 2011, 43 (10), 569-578.

(43)

Singh, N.; Ghosh, K. K.; Marek, J.; Kuca, K. Effect of Some Pyridinium Based Compounds on Hydrolysis of Carboxylate Ester. Indian J. Chem. 2012, 51 (B), 611−616.

(44)

Gupta B.; Sharma, R.; Singh, N.; Karpichev, Y.; Satnami, M. L.; Ghosh, K. K. Reactivity Studies Of Carbon, Phosphorus And Sulfur-Based Acyl Sites With Tertiary Oximes In Gemini Surfactants. J. Phys. Org. Chem. 2013, 26 (8), 632–642.

(45)

Ghosh, K. K.; Satnami, M. L. Nucleophilic Substitution Reaction Of Carboxylate And Phosphate Esters With Hydroxamate Ions In Microemulsions. Colloids Surf. A Physicochem. Eng. Asp. 2006, 274 (1-3), 125-129.

(46)

Marek, J.; Stodulka, P.; Cabal, J.; Soukup, O.; Pohanka, M.; Korabecny, J.; Musilek, K.; Kuca, K. Preparation of the Pyridinium Salts Differing in the Length of the N-Alkyl Substituent. Molecule, 2010, 15 (3), 1967-1972.

(47)

Albert, A.; Sergeant, E. P. Determinations of Ionization Constants, A Laboratory Manual; Chapman and Hall: London, 1971.

(48)

Um, I. H.; Shin, Y. H.; Lee, S. E.; Yang, K.; Buncel, E. Alkali Metal Ion Catalysis and Inhibition in Nucleophilic Displacement Reactions at Phosphorus Centers:  Ethyl and Methyl Paraoxon and Ethyl and Methyl Parathion. J. Org. Chem. 2008, 73 (3), 923-930.

34 ACS Paragon Plus Environment

Page 35 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(49)

Tarkka, R. M.; Buncel, E. Origin of the Bell-Shaped α-Effect−Solvent Composition Plots. pKa-Solvent Dependence of the Alpha-Effect at a Phosphorus Center. J. Am. Chem. Soc. 1995, 117 (5), 1503-1507.

(50)

Nagelkerke, R.; Thatcher, G. R. J.; Buncel, E. Alkali-Metal Ion Catalysis and Inhibition In Nucleophilic Displacement Reactions At Carbon, Phosphorus And Sulfur Centres. IX.P-Nitrophenyl Diphenyl Phosphate. Org. Biomol. Chem. 2003, 1, 163-167.

(51)

Terrier, F. ; Dafonte, P. R. ; Guevel, E. L. ; Moutiers, G. Revisiting The Reactivity Of Oximate Α-Nucleophiles With Electrophilic Phosphorus Centers. Relevance To Detoxification Of Sarin, Soman And DFP Under Mild Conditions. Org. Biomol. Chem. 2006, 4 (23), 4352-4362.

(52)

Popov, A. F. Design Of Green Microorganized Systems For Decontamination of Ecotoxicants. Pure Appl. Chem. 2008, 80 (7), 1381-1397.

(53)

Buncel, E.; I. H. Um, H. The Α-Effect And Its Modulation By Solvent. Tetrahedron 2004, 60 (36), 7801-7825.

(54)

Han, X.; Balakrishanan, V. K.; vanLoon, G. W.; Buncel, E. Degradation of the Pesticide Fenitrothion as Mediated by Cationic Surfactants and α-Nucleophilic Reagents. Langmuir 2006, 22 (21), 9009-9017.

(55)

Jurok, R.; Svobodova, E.; Cibulka, R.; Hampl, F. Reactivity in Micelles-are we Really Able to Design Micellar Catalysts? Collect. Czech. Chem. Commun. 2008, 73 (2), 127−146.

(56)

Simanenko, Y. S.; Karpichev, E. A.; Prokop’eva, T. M.; Panchenko, B. V. Micelles of an Oxime-Functionalized Imidazolium Surfactant. Reactivities at Phosphoryl and Sulfonyl Groups. Langmuir 2001, 17, 581−582.

(57)

Simanenko, Y. S.; Karpichev, E. A.; Prokopeva, T. M.; Lattes, A.; Popov, A. F.; Savelova, V. A.; Belousova, I. A. Functional Detergents Containing an Imidazole Ring and Typical Fragments of α- Nucleophiles Underlying Micellar Systems for Cleavage of Esters of Phosphorus Acid. Russ. J. Org. Chem. 2004, 40 (2), 206−218.

(58)

Simanenko, Y. S.; Popov, A. F.; Karpichev, E. A.; Prokopeva, T. M.; Savelova, V. A.; Bunton, C.A. Micelle Effects of Functional Surfactants, 1-Cetyl-3(2-Hydroxyiminopropyl)Imidazolium Halides, in Reactions with p-Nitrophenyl p-Toluene Sulfonate,

35 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 39

Diethyl p-Nitrophenyl Phosphate, and Ethyl p-Nitrophenyl Ethylphosphonate. Russ. J.Org. Chem. 2002, 38 (9), 1314−1325. (59)

Kapitanov, I. V.; Belousova, I. A.; Shumeiko, E.; Kostrikin, M. L.; Prokop’eva, T. M.; Popov, A. F. Supernucleophilic Systems Based On Functionalized Surfactants in The Decomposition Of 4-Nitrophenyl Esters Derived From Phosphorus And Sulfur Acids: II. Influence of The Length of Hydrophobic Alkyl Substituents On Micellar Effects Of Functionalized Monomeric And Dimeric Imidazolium Surfactants. Russ. J.Org. Chem. 2014, 50 (5), 694−704.

(60)

Bunton, C. A.; Ihara, Y. Micellar Effects Upon Dephosphorylation and Deacylation by Oximate Ions. J. Org. Chem. 1977, 42 (17), 2865-2869.

(61)

Couderc, S.; Toullec, J. Catalysis of Phosphate Triester Hydrolysis by Micelles of Hexadecyltrimethylammonium anti-Pyruvaldehyde 1-Oximate. Langmuir 2001, 17 (13), 3819-3828.

(62)

Biresaw, C.; Bunton, C. A. Size vs Reactivity in “Organized Assemblies”: Deacylation and Dephosphorylation in Functionalized Assemblies. J. Phys. Chem. 1986, 90, 5849−5853.

(63)

Bhattacharya, S.; Kumari, N. Metallomicelles as Potent Catalysts for the Ester Hydrolysis Reactions in Water. Coord. Chem. Rev. 2009, 253 (17-18), 2133−2149.

(64)

Andre, G. S.; Kliachyna, M.; Kodepelly, S.; Leriche, L. L.; Gillon, E.; Renard, P. Y.; Nachon, F.; Baati, R.; Wagner, A. Design, Synthesis and Evaluation of New ΑNucleophiles for The Hydrolysis of Organophosphorus Nerve Agents: Application To The Reactivation Of Phosphorylated Acetylcholinesterase. Tetrahedron 2011, 67, 63526361.

(65)

Bunton, C. A.; Nome, F.; Quina, F. H.; Romsted L. S. Ion Binding and Reactivity at Charged Aqueous Interfaces. Acc. Chem. Res., 1991, 24 (12), 357–364.

(66)

Silva, M.; Mello, R. S.; Farrukh, M. A.; Venturini, J.; Bunton, C. A.; Milagre, H. M. S.; Eberlin, M. N.; Fiedler, H. D.; Nome, F. The Mechanism of Dephosphorylation of Bis(2,4-dintrophenyl) phosphate in Mixed Micelles of Cationic Surfactants and Lauryl Hydroxamic Acids. J. Org. Chem. 2009, 74 (21), 8254−8260.

36 ACS Paragon Plus Environment

Page 37 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(67)

Tiwari, S.; Ghosh, K. K.; Marek, J.; Kuca, K. Functionalized Surfactant Mediated Reactions of Carboxylate, Phosphate and Sulphonate Esters. J. Phys. Org. Chem. 2010, 23 (6), 519−525.

(68)

Werawatganone, P.; Muangsiri, W. Interactions between Charged Dye Indicators and Micelles to Determine the Critical Micelle Concentration. Asian J. Pharm. Sci. 2009, 4 (4), 221−227.

(69)

Gonc, L. M.; Kobayakawa, A. T. G.; Zanette, D.; Chaimovichi, H.; Cuccovia, I. M. Effects of Micelles and Vesicles on the Oximolysis of p-Nitrophenyl Diphenyl Phosphate: A Model System for Surfactant- Based Skin-Defensive Formulations against Organophosphates. J. Pharm. Sci. 2009, 98 (3), 1040−10413.

(70)

Berezin, I. V.; Martinek, K.; Yatsimirskii, A. K. Physicochemical Foundations of Micellar Catalysis. Russ. Chem. Rev., 1973, 42 (10), 787–802.

(71)

Bunton, C. A.; Romsted, L.S.; Sepulveda. L. A Quantitative Treatment of Micellar Effects Upon Deprotonation Equilibria. J. Phys. Chem. 1980, 84 (20), 2611-2618.

(72)

Mchedlov-Petrossyan, N. O. Protolytic equilibrium in lyophilic nanosized dispersions: Differentiating influence of the pseudophase and salt effects. Pure. Appl. Chem. 2008, 80 (7), 1459-1510.

(73)

Kirby, A. J.; Nome, F. Fundamentals of Phosphate Transfer. Acc. Chem. Res. 2015, 48 (7), 1806–1814.

(74)

Epstein, J.; Kaminski, J. J.; Bodor, N.; Enever, R.; Sowa, J.; Higuchi T. Micellar acceleration of organophosphate hydrolysis by hydroximinomethylpyridinium type surfactants. J. Org. Chem., 1978, 43 (14), 2816–2821. 

(75)

Simanenko, Y.S.; Savelova, V.A.; Prokop'eva, T.M.; Mikhailov, V.A.; Turovskaya, M.K.;

Karpichev,

E.A.;

Popov,

A.F.;

Gillitt,

N.D.;

Bunton,

C.A.

Bis(dialkylamide)hydrogen Dibromobromate Precursors of Hypobromite Ion In Reactions With Nerve And Blister Agent Simulants. J. Org. Chem. 2004, 69 (26), 92389240. (76)

Prokop'eva, T. M.; Mikhailov, V.A.; Turovskaya, M.K.; Karpichev, E.A.; Burakov, N. I.; Savelova, V. A.; Kapitanov, I. V.; Popov, A. F. New sources of "active" halogen bis(dialkylamide)hydrogen dibromobromates, efficient reagents for destruction of ecotoxicants. Russ. J. Org. Chem. 2008, 44 (5), 637-646. 

37 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(77)

Page 38 of 39

Moss, R.A.; Kanamathareddy, S.; Vijayaraghava, S. Kinetics of Cleavage of Paraoxon and Parathion by Cetyltrimethylammonium Iodosobenzoate. Langmuir, 2001, 17(20), 6108–6112.

(78)

Turovskaya, M. K.; Kapitanov, I. V.; Belousova, I. A.; Tuchinskaya, K. K.; Shumeiko, A. E.; Kostrikin, M. L.; Razumova, N. G.; Prokop’eva, T. M.; Popov, A. F. Reactivity of Micelle-forming 1-alkyl-3-(1-oximinoethyl)pyridinium bromides in acyl group transfer reactions. Theor. Exp. Chem. 2011. 47 (1), 24 – 29

(79)

Kapitanov, I. V. Nucleophilicity of micellar systems based on amphiphilic derivatives of 2-(oximinomethyl)imidazole in the decomposition of 4-nitrophenyl diethyl phosphate. Theor. Exp. Chem. 2011, 47 (5), 317 – 323.

(80)

Jencks, W. P.; Carriuolo, J. Reactivity of nucleophilic reagents toward esters. J. Am.. Chem. Soc. 1960, 82 (7), 1778-1786.

(81)

Simanenko, Y.S.; Popov, A.F.; Prokop'eva, T.M.; Karpichev, E.A.; Savelova, V. A.; Suprun, I. P.; Bunton, C. A. Inorganic anionic oxygen-containing alpha-nucleophileseffective acyl group acceptors: Hydroxylamine ranks first among the alpha-nucleophile series. Russ. J. Org. Chem. 2002, 38 (9), 1286-1298.

(82)

Blanch, J. H.; Onsager, O.T. Stability of N-Heterocyclic Oxime Derivatives.Part II. Decomposition of N-Methylacetylpyridinium O-Acetylketoxime Iodides in Aqueous Solution. J. Chem. Soc. 1965, 3734-3738.

(83)

Blanch, J. H. Stability of N-heterocyclic oxime derivatives. Part III. The kinetics of the hydrolysis of formyl- and acetyl-pyridine O-acetyloximes in aqueous solution in the pH range 6·0–10·8 at 25, 35, and 40° J. Chem. Soc. B, 1968, 2, 167-169.

(84)

Kapitanov, I. V.; Abakumov, A. A.; Serdyuk, A. A. Identification of products in the reaction of 2-[(hydroxyimino) methyl]-1, 3-dimethylimidazolium iodide with diethyl 4nitrophenyl phosphate in alkaline medium. Russ.J. Org. Chem., 2015, 51 (10), 13681375.

38 ACS Paragon Plus Environment

Page 39 of 39

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 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

(85)

Blanch, J. H.; Andersen, J. Stability of N-Heterocyclic Oxime Derivatives. Part IV. Kinetics of Sodium Hydroxide-Catalysed Hydrolysis of Phosphonylated Ketoximes In Water at 15, 20, 25, and 30 Degrees. J. Chem. Soc. B, 1968, 2, 169-173.

(86)

Gorecki, L.; Korabecny, J.; Musilek, K.; Malinak, D.; Nepovimova, E.; Dolezal, R.; Jun, D.; Soukup, O.; Kuca, K. SAR study to find optimal cholinesterase reactivator against organophosphorous nerve agents and pesticides. Arch. Toxicol. 2016, doi: 10.1007/s00204-016-1827-3.

(87)

Bunton, C. A.; Hamed, F. H.; Romsted, L. S. Quantitative treatment of reaction rates in functional micelles and comicelles. J. Phys. Chem. 1982, 86 (11), 2103–2108.

(88)

Pehkonen, S.O.; Zhang, Q. The Degradation of Organophosphorus Pesticides in Natural Waters: A Critical Review. Crit. Rev. Environ. Sci. Technol. 2002, 32 (1), 17-72.

(89)

Druzina, B.; Stegu, M. Degradation Study of Selected Organophosphorus Insecticides In Natural Waters. Int. J. Environ. Anal. Chem. 2007 87 (15), 1079-1093

(90)

Dyguda-Kazimierowicz, E.; Roszak, Sz.; Sokalski, W.A. Alkaline Hydrolysis of Organophosphorus Pesticides: The Dependence of the Reaction Mechanism on the Incoming Group Conformation. J. Phys. Chem. B 2014, 118 (26), 7277-7289.

(91)

Liu, Y.; Zhang, C.; Liao, X.; Luo, Y.; Wu, S. Wang, J. Hydrolysis mechanism of methyl parathion evidenced by Q-Exactive mass spectrometry. Environ. Sci. Pollut. Res. 2015, 22 (24), 19747-19755.

(92)

Fenner, K.; Canonica, S.; Wackett, L.P.; Elsner, M. Evaluating Pesticide Degradation in the Environment: Blind Spots and Emerging Opportunities. Science 2013, 341 (6147), 752-758.

(93)

Ismail, D. A.; Ahmed, S. M. Ahmed, H. M.; Awad, A.I.; El-Sharkawy, H.A. Synthesis and Biological Activity of Alkyl Pyridinium Aldoxime Based Surfactants. Tenside Surfact. Det. 2016, 53 (4), 319-323.

39 ACS Paragon Plus Environment