l w; r (I If II I ll . w i 1' ‘ w Ir "H I y | M 7‘ ___ If I I 'H fl -H#-‘ Ioom (0—4—8 KINETJCS AND MECHANESM OF THE BERTHELOT REACTION WITH PARTECULAR REFERENCE '50 AMMONIA. ANALYSIS Thesis fer the Degree of M. S. MECHJGAN STATE UNWERSETY CHAREES JOHNSTON PAWON 3.975 lugma \J:M., ”,. V' -_.———q Qiigk- - “wofw gg* h‘fiq 7 ABSTRACT KINETICS AND MECHANISM OF THE BERTHELOT REACTION WITH PARTICULAR REFERENCE TO AMMONIA ANALYSIS By Charles Johnston Patton An investigation of the mechanism of the Berthelot reaction has led to an increased understanding of the many conflicts in the literature concerning the reaction‘s use for the analytical deter— mination of ammonia. Monochloramine is shown to be the first inter- nediate, and the kinetics of its formation and decomposition is pre- sented. The identical visible spectra obtained for a series of in- dophenols synthesized in the course of this research and the blue complexes formed when phenol or a substituted phenol was used in the Berthelot reaction has confirmed that indophenol is the final prod- uct of the reaction. Conditions which led to the formation of final products other than indophenol have been observed and are also dis- cussed. Finally, a new procedure for ammonia analysis using the Berthelot reaction, which is superior in both sensitivity and pre- cision to the best procedure found in the literature, is presented. KINETICS AND MECHANISM OF THE BERTHELOT REACTION WITH PARTICULAR REFERENCE TO AMMONIA ANALYSIS By Charles Johnston Patton A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemistry l975 DEDICATED TO LEROY N. HAYNES The blues ain't nothin' but a low-down shakin' chill. If you ain't never had 'em, God knows I hope you never will. Robert Johnson ii ACKNOWLEDGMENTS The author wishes to express his appreciation to Dr. S. R. Crouch for his guidance and encouragement during the course of this investigation. A large measure of thanks and gratitude are also due to Philip Notz, who spent many hours modifying and computer interfacing the stopped-flow spectrophotometer used in this research. Dr. Eric R. Johnson also deserves mention for writing sever- al computer programs used in this work. Many thanks to Dr. Peter Kissinger for his guidance in Dr. Crouch's absence; his stern, fatherly advice; and an endless supply of gin and tonics. Thanks also to his research group for late night comradship. The author also wishes to thank Brother Dave, Mary, his parents, and his parents-in-law for their support and encouragement. Finally, to all those individuals never mentioned in "Acknow- ledgment" sections, who made life a little less bearable by their ex- istance, kindly observe the mistletoe pinned to the tails of the author's coat as he walks down the corridors of the chemistry build- ing for the last time. 1°11 TABLE OF CONTENTS Page LIST OF TABLES ....................... vii LIST OF FIGURES ...................... viii I. INTRODUCTION ..................... 1 II. HISTORICAL ...................... 3 Evolution of the Berthelot Reaction ........ 3 Kinetics and Mechanism of the Berthelot Reaction . . 4 Role of Pentacyano Iron Complexes in the Berthelot Reaction ..................... 6 Use of Substituted Phenols in the Berthelot Reaction ..................... 8 III. EXPERIMENTAL ..................... 10 Equipment and Instruments ............. 10 Chemicals ..................... 10 Standardization Procedures ............. 10 Data Analysis ................... 11 IV. SYNTHESIS ...................... 12 Preparation of Indophenol and Related Compounds . . 13 Synthesis of 4-nitroso alkylphenols ...... 14 Synthesis of 4-nitroso chlorophenols ...... 15 2-methyl-4-nitrosophenol ............ 15 2,6-dimethy1-4-nitrosophenol .......... l6 3,5-dimethyl-4-nitrosophenol .......... 16 3-methyl-6-isopropy1-4-nitrosophenol ...... 16 2,3,5,6-tetramethy1-4-nitrosophenol ...... 16 2-chloro-4-nitrosophenol ............ l6 2,6-dichloro-4-nitrosophenol .......... 17 Synthesis of 5-nitroso-N-methy1anthranilic acid 17 Synthesis of 2-hydroxy-5-nitrosobenzoic acid . . 17 Synthesis of substituted 4-aminophenols . . . . 18 2-methy1-4-aminophenol ............. 18 2,6-dimethy1-4-aminophenol ........... 18 3,5-dimethyl-4-aminophenol ........... 19 3-methy1-6-isopropy1-4-aminophenol ....... 19 2,3,5,6-tetramethyl-4-aminophenol ....... 19 2-chloro-4-aminophenol ............. 19 iv 2,6-dichloro-4-aminophenol ........... 2-hydroxy-5-amino benzoic acid ......... Synthesis of quinone-N-chloroimines ...... Quinone-N-chloroimine ............. Z-methylquinone-N-chloroimine ......... 2,6-dimethquuinone-N-chloroimine ....... 3,5-dimethy1quinone-N-chloroimine ....... 3-methy1-6-isopropy1quinone-N-chloroimine 2,3,5,6-tetramethylquinone-N-chloroimine . . . . 2-chloroquinone-N-chloroimine ......... 2,6-dichloroquinone-N-chloroimine ....... Synthesis of indophenols ............ Indophenol, sodium salt ............ 2,2'-dimethylindophenol, sodium salt ...... 2,6,2',6'-tetramethylindophenol, sodium salt . . 3,3'-dimethyl-6,6'-diisopropy1indophenol, sodium salt ...................... 3,5,3',5'-tetramethy1indophenol, sodium salt . . 2,3,5,6,2',3',5',6'-octamethylindophenol, sodium salt ..................... 2,2'-dich10roindophenol, sodium salt ...... 2,6,2',6'-tetrachloroindophenol, sodium salt . . Discussion ................... Preparation of Other Materials ........... Synthesis of sodium hypochlorite ........ Synthesis of monochloramine .......... Synthesis of sodium aquopentacyanoferrate Discussion ................... V. NON-KINETICS EXPERIMENTS ............... UV Spectrum of Hypochlorous Acid .......... UV Spectrum of Monochloramine ........... Visible Spectra of Pentacyano Iron Complexes . . . . Visible Spectra of Indophenols ........... Formation of Indophenols in Aqueous Solution . . . . Identity of the "Catalyst" for the Berthelot Reaction ..................... VI. KINETICS EXPERIMENTS ................. Formation of Monochloramine ............ Decomposition of Monochloramine .......... Mechanism and Kinetics of Indophenol Formation . Rates of Formation of Several Indophenols ..... VII. A NEW AMMONIA METHOD ................. Reagents ...................... Ammonia Standards ............... Sodium hypochlorite .............. Sodium aquopentacyanoferrate ...... ‘. . . . 2-chlorophenol ................. The Method ..................... Discussion ................... LIST OF REFERENCES ..................... vi LIST OF TABLES Table Page 1. Absorption Maxima and Molar Absorptivities of Some Pentacyano Iron Complexes ........ 32 2. Absorption Maxima and Molar Absorptivities of Some Indophenols and Their Sodium Salts . . . . 34 3. Effect of Aquopentacyanoferrate Concentration on the End Product of the Berthelot Reaction . . . 36 4. Effect of Aquopentacyanoferrate Concentration on the End Product of the Berthelot Reaction in the Presence of Hypochlorite Ions ....... 36 5. Data Obtained When Ammonia Was Analyzed by the Proposed Method ................. 56 6. Data Obtained When Ammonia Was Analyzed by the Solorzano Method ................. 57 vii Figure 10. 11. 12. 13. LIST OF FIGURES UV Spectrum of Hypochlorous Acid as a Function of pH ...................... UV Spectrum of Hypochlorite Ions ......... UV Spectrum of Monochloramine ........... Visible Spectrum of 2,6,2',6'-tetramethy1indophenol as a Function of pH ............... Formation of Monochloramine from Ammonia and Hypo- chlorous Acid in 1:1 Molar Ratio ........ Formation of Monochloramine from Ammonia and Hypo- chlorous Acid in 1:5 Molar Ratio ........ Formation of Monochloramine from Ammonia and Hypo- chlorous Acid in 1:10 Molar Ratio ........ Log of Observed Second-order Rate Constants for the Formation of Monochloramine versus pH at 292 nm (o) and 243 nm (I) ............ Distribution Diagrams for Hypochlorous Acid- Hypochlorite Ions and Ammonia-Ammonium Ions as a Function of pH ............... Decomposition of Monochloramine as a Function of Absorbance at 243 nm and Time at pH values of 8.5 (I), 10.5 (0), and 12.0 (A) ......... Decomposition of Monochloramine as a Function of Absorbance at 292 nm and Time at pH Values of 8.5 (I), 10.5 (o), and 12.0 (A) ......... Rate of Formation of 2,2'-dimethy1indophenol (A), Indophenol (I), and 2,2'-chloroindophenol (0) . . Calibration Curves for Ammonia as Determined by the Solorzano (o) and Proposed (I) Methods viii Page 30 31 31 33 41 42 43 44 45 48 48 50 58 I. INTRODUCTION Although a large number of analytical schemes for ammonia analysis have appeared in the literature, the method finding the most widespread application is the Berthelot reaction. In this reaction, indophenol, an intensely blue charge-transfer complex, is formed when an aqueous solution containing ammonia is treated with phenol and hypochlorous acid. Under controlled conditions, the amount of indophenol produced is proportional to the concentration of ammonia originally present in the solution. The method is specific for am- monia, and when sodium nitroprusside is used as a "catalyst," detec- tion limits for ammonia concentrations as low at 10'7 M_have been reported. The precision of such procedures is quite poor, however, and despite the many papers which have appeared in the literature describing "improved" procedures for the catalyzed Berthelot reac- tion, a truly satisfactory method has yet to appear. The trouble with most of the work to date on the Berthelot reaction is that while reaction parameters have been wildly juggled to "optimize" various procedures, little attention has been paid to the mechanism which leads to indophenol formation. In this research, a mechanistic approach to the problem of poor precision in the Ber- thelot reaction was undertaken. The rate of formation of monochlor- amine, the first intermediate in the Berthelot reaction, was deter- mined as a function of pH. The rate was maximum at pH 8.0 with an 4 3 1 observed second—order rate constant of 2.4 x 10 i 0.4 x 10 a mole- sec-1. It was also observed that dilute aqueous solutions of mono- chloramine decomposed extensively in the presence of hypochlorous acid if the pH was less than 10.0. In other experiments, the actual catalyst for the Berthelot reaction was found to be sodium pentacyanoaquoferrate, rather than sodium pentacyanonitroferrate or sodium pentacyanonitritoferrate as reported by other workers. This observation has not been reported elsewhere. Also, a series of substituted indophenols was synthesized, and the details of their preparation, as well as their visible spec- tra, have been included. These data confirmed that indophenol was the final product of the Berthelot reaction. The final outcome of this research was a new analytical pro- cedure for ammonia which has a lower detection limit and higher pre- cision than any published procedure for the catalyzed Berthelot re- action. This method could be easily automated and, with minor modifi- cations, should be well-suited to the analysis of ammonia in natural waters. II. HISTORICAL Evolution of the Berthelot Reaction The blue color produced in alkaline solutions containing phenol, hypochlorite ions, and ammonia was first reported by Berthe- lot (4). This reaction was first used analytically by Thomas (42) to determine ammonia in blood. Although several other workers (6,43) modified Thomas's procedure, the first real improvement in ammonia analysis via the Berthelot reaction was reported by Russel (35). She found that the reaction was catalyzed by Mn(II) ions, and ob- served a three-fold increase in sensitivity relative to the uncata- lyzed reaction. Ten years later, two other catalysts more effective than Mn(II) ions were reported: acetone (10) and sodium nitroferricya- nide (22). The use of these new catalysts, especially sodium nitro- pentacyanoferrate (nitroprusside), increased the sensitivity of the Berthelot reaction to the extent that concentration of the ammonia in biological fluids and natural waters prior to analysis was no longer necessary. This, coupled with the reaction's relative insen- sitivity to nitrogen-containing compounds other than ammonia, led to its widespread application. Over the past twenty years, literally scores of papers have appeared describing new applications or modi- fied procedures for the catalyzed Berthelot reaction. Unfortunately, few of these papers agree on such reaction parameters as reagent concentrations, pH, or order and timing of reagent additions. For example, a few papers (5,36) reported en- hanced sensitivity when the chlorinating reagent was added first, while most other papers reported virtually no sensitivity unless the phenol reagent was added first. This led some researchers (15,37) to the erroneous conclusion that the phenol-hypochlorite and hypochlorite-phenol reactions were two distinct entities. Numer- ous conflicts in the literature give the impression that no single mechanism is operative in the catalyzed Berthelot reaction since the only thing most procedures have in common is extremely poor preci- sion. Kinetics and Mechanism of the Berthelot Reaction Very little has been published concerning the kinetics and mechanism of the Berthelot reaction. Russel (35) suggested that the end product of this reaction was indophenol or a closely related com- pound. Bolleter, gt;_al,, (5) proposed a generalized mechanism (out- lined below) in which monochloramine and quinone-N-chloroimine were intermediates and indophenol was the end product. The role of a catalyst was not considered. _ 11> N33 + on L—s» H2NCI + ‘OH monochloroimine H2NCI + .an 4911—» o=.=Nc1 quinone-N-chloroimine a... + cm‘o JO—H. oHN 0’ NJ indophenol, sodium salt There is substantial evidence in the literature (7,11,24) which supports the first step of Bolleter's proposed mechanism. At pH values greater than 8.5, monochloramine is the sole product in the reaction between equimolar concentrations of ammonia and hypo- chlorite ions. The fact that most papers in which reaction param- eters for the Berthelot reaction were investigated reported a drastic reduction in sensitivity when the hypochlorite reagent was added first, however, led many workers to discount the proposed mechanism. This conflict over the preferred order of reagent addition seems to have been resolved by Horn and Squire (18). They noted that all procedures in which the phenol reagent had to be added first in order to obtain satisfactory sensitivity used hypochlorite reagents in at least one thousand-fold excess of the stoichiometric amount necessary to convert all the ammonia in the sample to monochloramine. They found, however, that the order of reagent addition could be re- versed, affecting an increase in sensitivity and precision in these procedures if the concentration of the hypochlorite reagent was re- duced by two orders of magnitude. This observation suggested that monochloramine was unstable in the presence of hypochlorite ions. It is interesting to note that Morris and Neil (27), who did exten- sive research on the products formed when waste water was chlorinated, reported that monochloramine was unstable in solutions containing a large excess of hypochlorite ions. Role of Pentacyano Iron Complexes in the Berthelot Reaction In a subsequent paper (19), Horn and Squire presented evi- dence that the catalyst for the Berthelot reaction was sodium nitri- topentacyanoferrate which formed spontaneously when sodium nitro- ferricyanide, the assumed catalyst, was prepared in mildly alkaline solutions. They reported that conversion of the nitro complex to the nitrito complex prior to its use as a reagent increased both the sensitivity and precision of the Berthelot reaction. Later, Mark (23) confirmed this observation and suggested that the nitrito com- plex catalyzed the formation of monochloramine. There is evidence in the literature not pertaining directly to the Berthelot reaction, however, which suggests that Mark's hy- pothesis concerning the catalytic role of sodium nitritopentacyano- ferrate is in error. First, work of Neil and Morris (47) and of Yagil and Anbar (49) has shown that the formation of monochloramine from ammonia and hypochlorous acid in the pH range of 7.5 to 11.5 is too rapid to be studied by conventional techniques. It does not seem likely that a reaction this rapid would require a catalyst. Second, work of Ohkuma (30) and of Swinehart and Rock (41) has shown that while sodium nitropentacyanoferrate is converted to sodium nitrito- pentacyanoferrate, the nitrito complex is in equilibrium with sodium aquopentacyanoferrate, as shown in the following scheme. For this [Fe(CN)5NO]2- + 2'0H sé [Fe(CN)5N02]4' + H20 [Fe(CN)5N02]4' + H20 # [Fe(CN)5H20]3' + N02- reason, it is possible that the real catalyst for the Berthelot re- action is the aquo complex rather than either the nitro or the ni- trito complexes. In another paper (29), Ohkuma reported that sodium nitrofer- ricyanide is converted directly to sodium aquopentacyanoferrate by alkaline hydroxylamine. The aquoferrate thus formed incorporates a second molecule of hydroxylamine as a ligand, and in the presence of phenol, aminophenol was produced as shown below. The similarity be- Na2[Fe(CN)5NO] + HZNOH + NaOH -———) Na3[Fe(CN)5H20] + N2 + -OH ' 'OH Na3[Fe(CN)5H20] + HZNOH ——> Na3[Fe(CN)5H2NOH] + H20 Na3[Fe(CN)5H2NOH] + .ou -°—”+ HO—.—NH2 + Na3[Fe(CN)5H20] tween monochloramine and hydroxylamine suggests first that an analogous mechanism may explain the catalytic role of sodium nitro- ferricyanide in the Berthelot reaction, and second that sodium aquo- pentacyanoferrate may be a better catalyst than either sodium nitro- pentacyanoferrate or sodium nitritopentacyanoferrate. To date, there has been no mention in the literature concerning the relevance of these works to the elucidation of the reaction mechanism for the catalyzed Berthelot reaction. Use of Substituted Phenols in the Berthelot Reaction Another interesting development in the evolution of the Ber- thelot reaction as an analytical tool is the use of substituted phenols. Roskam and de Langen (34) reported a two-fold increase in sensitivity and more stable blanks when thymol (3-methyl-6- isopropylphenol) was used in place of phenol in the acetone- catalyzed Berthelot reaction. Procedures in which salicylic acid (33), 2-chlorophenol (51), 3-methy1phenol (50), and l-naphthol (26) were used have also been reported. Although no mention was made in these references concerning the mechanistic or kinetics implications of using substituted phenols in the Berthelot reaction, the work of several dye chemists is relevant. Corbett (9) reported a synthesis for indophenols in solution as sodium salts by the ferricyanide oxidation of aminophenols to the corresponding quinone imines. The imines thus formed coupled spon- taneously with phenol to give indophenolate ions when the pH range was between 8.5 and 11.5, as shown in the following scheme. Corbett 2[Fe(CN) 613' + H0—.—NH2 T—-———* O-.=NH + 2[Fe(CN) 614' + 2H+ quinone imine O=.=NH + .OH :91» Na+'O—.—N‘.=O found that in the preparation of methyl substituted indophenolate ions, the use of the less methylated aminophenols and the more meth- ylated phenols led to the fastest coupling rates. Thus when quinone monoimine was reacted with 2- and 3-methylphenol, for example, there was a thirty- and ten-fold increase, respectively, in the rate of in- dophenolate ion formation relative to phenol. In the case of chloro- indophenolate ion synthesis, the use of the more chlorinated amino- phenol and the less chlorinated phenol led to the fastest coupling rates. Gibbs, Hall, and Clark (14), in a much earlier work, made this same observation regarding the coupling rates of quinone-N- chloroimines and phenols although no quantitative data were given. O‘NCl + H‘OH + 20H —) OHNO- These observations suggest that there might be a kinetics advantage in using certain substituted phenols in the Berthelot reaction. III. EXPERIMENTAL Equipment and Instruments Melting point data were obtained with an oil bath immersion apparatus (Thomas-Ashley). A11 pH data were obtained with a servo digital pH/volt meter (Heath EU-302A) and a combination reference and pH electrode (Beckman). All visible and ultraviolet spectra were obtained with a scanning spectrophotometer (Cary 17). Infrared spectra were obtained with a scanning infrared spectrophotometer (Beckman AccuLab 2). Data from kinetics experiments were obtained by using an automated stopped-flow spectrophotometer designed in these laboratories (2,3). Chemicals The phenols used as starting materials for indophenol syn- thesis were obtained from three sources: Aldrich, Eastman Kodak, and K and K. All were marketed as = 97% purity and were used as re- ceived. All other chemicals used in this research were reagent grade, and no attempt was made to purify them further. Standardization Procedures Solutions of sodium hypochlorite and monochloramine were standardized by thiosulfate titration. Two grams of potassium 10 ll iodide and 5 ml of glacial acetic acid were added to an aliquot of the solution to be analyzed. The sample was immediately titrated to a starch end point with a 0.1 N_sodium thiosulfate, which had been previously standardized against primary standard grade potassium iodate. Phosphate and borax buffers in the pH range of 4.0 to 12.5 were prepared as described in the "Handbook of Chemistry and Physics" (46). These were referenced relative to commercial standard buffers of pH 4.00, 7.00, and 10.00. Data Analysis General calculations, including standard deviation and linear regression analysis, were performed on an SR-Sl calculator (Texas Instruments). Control of the stopped-flow spectrophotometer, data acquisition, and formating was performed by a POP/8e minicom- puter (Digital Equipment Corporation). Data from kinetics runs were analyzed with the aid of KINFIT, a generalized curve-fitting routine developed by Dye and Nicely (13). IV. SYNTHESIS Many of the compounds and reagents used in this research were not commercially available. The details of their preparation are included in this chapter. The first section covers the synthe- sis of symmetric indophenols, where the word symmetric implies that the identity and position of substituents are identical on both rings of the indophenol. The numbering system used for naming indophenols is shown below. The second section covers the synthesis of sodium hypochlorite, monochloramine, and sodium aquopentacyanoferrate, which are participants in the Berthelot reaction. All yields reported for compounds synthesized were calculated on the basis of total dry weight of the crude product recovered. Generally, only a small portion of the crude product was recrystal- lized or sublimed for melting point and spectra determination, while the rest was used as recovered for succeeding syntheses. Any signi- ficant details for a particular synthesis not included in the general 12 13 procedures for each class of compounds can be found at the end of each section. Preparation of Indophenol and Related Compounds Since the end product of the Berthelot reaction is generally assumed to be an indophenol dye, a series of these compounds was syn- thesized. TWo pathways for indophenol synthesis were investigated: the Liebermann nitroso reaction and the procedure outlined by Gibbs, 92.31.. In the Liebermann nitroso reaction, a phenol is reacted with nitrous acid in the cold to produce the nitrosophenol. In the pre- sence of excess phenol and strong mineral acid, the nitrosophenol and phenol couple to form the acid sulfate of indophenol, as shown below. Dilution with cold water precipitates the free indophenol. ‘+HONO —°°-> ‘e—S *3 14°44 ”HM: m,- 14 In the Gibbs procedure, a phenol is reacted with a quinone- N—chloroimine in basic media to form an indophenol as the sodium salt. + O 2NoOlL; _C N... NJ+N¢CI+2H20 Since the sodium salts of indophenols were found to be much more stable than the free indophenols from the Liebermann nitroso re- action, the Gibbs procedure proved to be the most satisfactory method for indophenol synthesis. In general, the 4-aminophenols necessary as starting material for indophenol synthesis were not available, and so a series of com- mercially available substituted phenols were 4-nitrosated and then reduced to the desired 4-amino compounds. Nitration of phenols and subsequent reduction to the amine proved less satisfactory for two reasons. First, nitration gives several products which must be se- parated, while nitrosation yields the 4-substituted products almost exclusively. Second, the nitrosophenols are reduced under much milder conditions than the corresponding nitrophenols. Synthesis of 4-nitroso alkylphenols For synthesis of 4-nitroso alkylphenols, 0.20 mole of the alkylphenol was added to a 400 m1 beaker containing 50 ml of ethanol and 50 ml of concentrated hydrochloric acid. An off-white slurry 15 formed after the reaction mixture was cooled to -2°C in a salt-ice bath. Next, 0.25 mole of pulverized sodium nitrite was added grad- ually while the reaction mixture was hand stirred with a glass rod. After about half of the sodium nitrite was added, the reaction mix- ture thickened considerably and a yellow-green sheen was observed on the sides of the reaction vessel. Fifteen minutes after the ad- dition of sodium nitrite was complete, the reaction mixture was poured into a l e beaker containing 700 ml of ice cold water. A co- pious, fluffy precipitate developed, which was suction filtered and washed well with cold water. Synthesis of 4-nitroso chlorophenols A 500 ml beaker containing a solution of 0.03 mole of the chlorophenol, 2 g of sodium hydroxide, and 10 g of sodium nitrite in 300 ml of water was cooled to 15°C in an ice bath. Next, 25 ml of 33% v/v sulfuric acid, cooled to 4°C, were added gradually. When the addition of sulfuric acid was complete, the reaction mixture was transferred to a refrigerator. After two hours, a large mass of burnt-orange needles had formed, which were filtered and washed with a little ice cold water. The mother liquor was returned to the re- frigerator, and after eight hours, a second crop of needles was re- covered. 2-metbyl-4-nitrosgphenol The reaction of 22.0 g (0.20 mole) of 2-methylphenol gave 22.0 g (80.2%) of the crude nitroso-phenol. Recrystallization from water gave white needles, MP = l34-135°C-d [lit (12):l34-135°C-d]. 16 2,6-dimethyl-4-nitrosophenol The reaction of 24.5 g (0.20 mole) of 2,6-dimethylphenol gave 18.7 g (61.8%) of the crude nitrosophenol. Recrystallization from benzene gave gold plates, MP = 17l-l73°C-d [lit (44): 171°C-d]. 3,5-dimetby1-4-nitrosophenol The reaction of 10.0 g (0.08 mole) of 3,5-dimethylphenol gave 6.5 g (53.7%) of the crude nitrosophenol. Recrystallization from benzene gave yellow plates, MP = 184-185°C-d [lit (28): 185-186°C-d]. 3-metbyl-6-isopropy1:4-nitrosophenol The reaction of 10.0 g (0.07 mole) of 3-methyl-6- isopropylphenol gave 8.6 g (68.6%) of the crude nitrosophenol. Re- crystallization from benzene gave white needles, MP = 162-163°C-d [lit (12): l60-164°C-d]. 2,3,5,6-tetramethyl-4-nitrosgphenol The reaction of 30.0 g (0.20 mole) 2,3,5,6-tetramethy1phenol gave 27.2 g (75.9%) of the nitrosophenol. Recrystallization from 50% ethanol gave yellow needles, MP = 88-89°C-d [lit (40): 89°C-d]. 2-chloro-4-nitrosophenol The reaction of 4.0 g (0.03 mole) of 2-chlorophenol gave 2.5 g (52.9%) of the nitrosophenol. Recrystallization from 50% etha- nol gave burnt-orange needles, MP = 147°C-d [lit (16): 148°C-d]. 17 2,6—dichloro-4-nitrosgphenol ‘ The reaction of 4.0 g (0.02 mole) of 2,6-dichlorophenol gave 2.7 g (57.4%) of the crude nitrosophenol. Recrystallization from benzene gave orange needles, MP = 162-164°C-d [lit (8): 160-164°C-d]. Synthesis of 5-nitroso-N-methylanthranilic acid N—methylanthranilic acid (0.07 mole) was dissolved in 75 m1 of concentrated hydrochloric acid. The solution was cooled to 0°C in a salt-ice bath, and 5 g of sodium nitrite were added gradually. The reaction mixture was transferred to a refrigerator, and after twelve hours the hydrochloride of 5-nitroso-N-methalanthranilic acid had separated as a bright gold precipitate, which was filtered and washed successively with 5 ml of cold ethanol and 40 m1 of ether. The hydrochloride was transformed into the free base by trituration with 0.05 mole of sodium bicarbonate and 300 ml of water. The blue- green precipitate which formed was filtered and dried. The yield was 9.5 g (75.4%). The crude product was recrystallized from 50% ethanol to give emerald green crystals, MP = l9l-192°C-d [lit (20): 191°C-d]. Synthesis of 2-hydroxy-5-nitrosobenzoic acid Crude S-nitroso-N-methylanthranilac acid (0.025 mole) was added to 25 ml ice cold 10 N_sodium hydroxide. To this 150 m1 of ice cold water were then added slowly. The temperature of the reaction mixture was always less than 5°C. After twelve hours in a refrigera- tor at 0°C, the reaction mixture had a deep redish-blue color. The 18 temperature was maintained at 0°C in a salt—ice bath, and ice cold concentrated sulfuric acid was added drop-wise until a pH of 1 was obtained. The blue-green precipitate which formed was filtered, washed with several small portions of ice cold water, and dried in a vacuum desiccator. The yield was 4.1 g (98.1%). The crude product was recrystallized from benzene to give blue-green needles, MP = 161-162°C—d [lit (20): 162-163°C-d]. Synthesis of substituted 4-aminophenols The 4-nitrosophenol (0.1 mole) was dissolved in 250 ml of ammonium hydroxide solution prepared from 160 ml of water and 90 ml of concentrated ammonium hydroxide. The deep yellow-brown solution was suction filtered and transferred to a 400 ml beaker. Next, tank hydrogen sulfide diffused through a sintered glass frit was vigorous- ly passed through the solution. After five minutes, the reaction mixture turned clear orange and a precipitate began to form. After twenty minutes, the precipitate was copious and the hydrogen sulfide was turned off. The reaction mixture was cooled to 0°C and filtered. 2-methyl-4-aminophenol The reaction of 10 g (0.07 mole) of crude 2-methyl-4- nitrosophenol gave 6.4 g (71.3%) of the amine. Recrystallization from benzene gave white plates, MP = 174-176°C [lit (12): 175°C]. 2,6-dimethyl—4-aminophenol The reaction of 5 g (0.03 mole) of recrystallized 2,6- dimethyl-4—nitrosophenol gave 3 g (66.3%) of the amine. 19 Recrystallization from benzene gave tan needles, MP = 138°C-d [lit (44): 137-138°C-d]. 3,5-dimethyl-4-aminophenol The reaction of 6 g (0.04 mole) of crude 3,5-dimethyl-4- nitrosophenol gave 5 g (91.1%) of the amine. Recrystallization from benzene gave white plates, MP = 182-183°C—d [lit (44): 181.8- 183°C-d]. 3-methyl-6-isopropy1-4-amingphenol The reaction of 9 g (0.05 mole) of crude 3-methy1-6- isopropyl-4-nitrosophenol gave 8 g (96.8%) of the amine. Recrystal- lization from benzene gave white needles, MP = l78-l79°C [lit (12): 178-179°C]. 2,3,5,6—tetramethyl-4-amin9phenol The reaction of 6 g (0.03 mole) of crude 2,3,5,6-tetramethyl- 4-nitrosophenol gave 4.7 g (86.2%) of the amine. Sublimation gave a white powder, MP = 179°C-d [lit (38): 177-178.5°C-d]. 2-chloro-4-aminophenol The reaction of 5.1 g (0.03 mole) of crude 2-chloro-4- nitrOSOphenol gave 2.9 g (67.3%) of the amine. Recrystallization from benzene gave white needles, MP = 151°C-d [lit (12): 153°C-d]. 2,6-dichloro-4-aminophenol The reaction of 2.4 g (0.01 mole) of crude 2,6-dichloro-4- nitrosophenol gave 1.6 g (71.9%) of the amine. Recrystallization 20 from benzene gave tan needles, MP = l68-l70°C-d [lit (12): 165- 166°C-d]. 2-hydroxyr5-amino benzoic acid When 2.3 g (0.02 mole) of 2-hydroxy-5-nitroso benzoic acid was reacted, no product was isolated. Synthesis of quinone-N-chloroimines The aminophenol to be reacted was dissolved in 100 m1 of 5% v/v hydrochloric acid (2 0.6 N) and cooled to )°C. This solution was then added gradually to a three- to five-fold excess of sodium hypochlorite solution also cooled to 0°C and containing a few pieces of crushed ice. A bright yellow to orange precipitate formed im- mediately. After five minutes, this precipitate was filtered, washed with cold water, and dried in a vacuum desiccator. Quinone-N-chloroimine The reaction of 5.5 g (0.05 mole) of commercial 4- aminophenol gave 5.7 g (81.4%) of the N-chloroimine. Recrystalliza- tion from ether gave yellow needles, MP = 87-89°C-d [lit (45): 88°C-d]. 2-methy1quinone-N-chloroimine The reaction of 1.2 g (0.01 mole) of 2-methyl-4-aminophenol gave 1.4 g (90%) of the N-chloroimine. Recrystallization from ether gave yellow needles, MP = 99-100°C-d [litz no data]. 21 2,6-dimethylquinone-N-chloroimine The reaction of 2.4 g (0.2 mole) of 2,6-dimethy1-4- aminophenol gave 2.1 g (61.9%) of the N-chloroimine. Recrystalliza- tion from benzene gave bright yellow needles, MP = 69.5-7l°C-d [litz no data]. 3,5-dimethquuinone-N-ch1oroimine The reaction of 4.6 g (0.03 mole) of 3,5-dimethyl-4- aminophenol gave 3.9 g (69.7%) of the N-chloroimine. Recrystalliza- tion from ether gave yellow needles, MP = 175-177°C-d [litz no data]. 3-methyl~6-isopropylquinone-N-chloroimine The reaction of 5.0 g (0.03 mole) of 3-methy1-6-isopropy1-4- aminophenol gave a bright orange oil. Repeated attempts to crystal- lize the product failed. 2,3,5,6-tetramethylquinone-N-chloroimine The reaction of l g (0.06 mole) of 2,3,5,6-tetramethyl-4- aminophenol gave 0.7 g (59.0%) of the N-chloroimine. The product ap- peared dirty and no suitable recrystallization solvent was found, MP = 77-80°C-d [litz no data]. 2-chloroquinone-N-chloroimine The reaction of 5.5 g (0.04 mole) of 2-chloro-4-aminophenol gave 3.5 g (52.3%) of the N-chloroimine. Recrystallization from ether gave orange needles, MP = 85-86°C-d [lit (24): 86-87°C-d]. 22 2,6-dichloroquinone-N-chloroimine The reaction of 8 g (0.04 mole) of commercial 2,6-dichloro- 4-aminophenol gave 3.3 g (35.2%) of the N-chloroimine. Recrystal- lization from ether gave gold needles, MP = 64-45°C-d [lit (25): 66-67°C-d]. Synthesis of indophenols A 100 ml beaker containing 30 ml of water and a few pieces of ice was cooled to 0°C in a salt-ice bath. Next equimolar quanti- ties of phenol and quinone-N-chloroimine were added. Then enough 3 N_sodium hydroxide to provide a 2:1 molar ratio plus 10% was added slowly while the temperature of the reaction mixture was maintained at 0°C. The reaction mixture immediately darkened and passed through several shades of green before finally attaining a deep blue color. The calculated amount of sodium chloride necessary to give a near saturated solution was added, and the reaction mixture was transferred to a refrigerator. After approximately four hours, the irradescent sodium salt of the indophenol had precipitated. The precipitate was suction filtered and dried in a vacuum desiccator over calcium chloride. Indophenol, sodium salt The reaction of 4.7 g (0.05 mole) of phenol, 7.1 g (0.05 mole) of quinone-N-chloroimine, and 38 ml of 3 N_sodium hydroxide solution gave 12.2 9 (109.4%) of the crude indophenol salt. 23 2,2'-dimethylindophenol, sodium salt The reaction of 1.0 g (0.009 mole) of 2-methylphenol, 1.4 g (0.009 mole) of 2-methquuinone-N-chlorimine, and 7 m1 of 3 N_sodium hydroxide solution gave 3.5 9 (156.1%) of the crude indophenol salt. 2,6,21,6'-tetramethylindgphenol,,sodium salt The reaction of 1.4 g (0.012 mole) of 2,6-dimethy1phenol, 2.0 g (0.012 mole) of 2,6-dimethquuinone-N-ch1oroimine, and 8 m1 of 3 N_sodium hydroxide solution gave 3.6 g (108.3%) of the crude in- dophenol‘salt. 3,3'-dimethyl-6,6'-diisgpropylindophenol, sodium salt The reaction of 0.8 g (0.005 mole) of thymol, 1.0 g (0.005 mole) of the orange oil thought to be 3-methyl-6-isopropy1quinone- N-chloroimine, and 4 ml of 3 N_sodium hydroxide solution gave a sticky, purple solid which could not be further purified. 3,5,3',5'-tetramethylindophenol,_sodium salt When 1.4 g (0.012 mole) of 3,5-dimethylphenol, 2.0 g (0.012 mole) of 3,5-dimethquuinone-N-ch1oroimine and 8 ml of 3 N_sodium hydroxide were reacted, the charactersitic blue color did not de- velop and no product was recovered. 2,3,5,6,2',3',5',6'-octamethylindophenol When 0.46 g (0.003 mole) of 2,3,5,6-tetramethylphenol, 0.60 g (0.003 mole) of the crude 2,3,5,6-tetramethquuinone-N-chloroimine and 2 ml of 3 N_sodium hydroxide solution were reacted, the charac- teristic blue color did not develop and no product was recovered. 24 2,2'-dichloroindgphenol, sodium salt The reaction of 2.3 g (0.018 mole) of 2-chlorophenol, 3.2 g (0.018 mole) of 2-chloroquinone-N-chloroimine, and 13 ml of 3 N_so- dium hydroxide solution gave 6.1 9 (116.8%) of the crude indophenol salt. 2,6,2',6'-tetrachloroindophenol, sodium salt The reaction of 2.8 g (0.013 mole) of 2,6-dichloroquinone-N- chloroimine and 8.5 m1 of 3 N_sodium hydroxide solution gave 4.9 g (105.10%) of the crude indophenol salt. Discussion 4-nitroso alkylphenols and 4-aminophenols were prepared by the Kremers and Wakeman procedure (31), and 4-nitroso chlorophenols by the Hodgson procedure (16). The synthesis of 5-nitroso-N- methylanthranilic acid and 2-hydroxy-5-nitrosobenzoic acid followed the procedure of Houben (20). Quinone-N-chloroimines and sodium salts of indophenols were prepared by the procedures of Gibbs, gt, 21, (14). Generally, reduction of nitrosophenols to aminophenols with hydrogen sulfide proceded smoothly. It proved essential, however, to filter the ammoniacal nitrosophenol solutions prior to hydrogen sulfide reduction in order to obtain the aminophenols in high purity. Due to the insolubility of 2,3,5,6-tetramethyl#4-nitrosophenol in water, it was necessary to dilute the concentrated ammonium hydrox- ide with ethanol rather than with water. After this reaction was 25 complete, addition of 50 ml of water enhanced the precipitation of the aminophenol. The oxidation of aminophenols to quinone-N-chloroimines also proceded smoothly. Precooling both the hypochlorite and acidic aminophenol solutions to 0°C and maintaining that temperature in the reaction mixture was essential. The fact that some of the amino- phenols were not completely soluble in 5% v/v hydrochloric acid so- lution did not seem to hinder their complete reaction. For all 4-amino alkylphenols reacted, a commercial sodium hypochlorite solu- tion, ChloroxR, was a satisfactory reagent. For 4-amino chloro- phenols, however, Rashig's hypochlorite, prepared as described in the next section, was a better reagent. The difficulty in synthesizing symmetrical indophenols arose in their separation and purification. In all cases, sodium chloride had to be added to "salt out" the product, and invariably the sodium salts of indophenols thus obtained were contaminated with sodium chloride. This accounts for the greater than 100% yields reported for the indophenols synthesized. Two purification procedures were investigated. In one procedure, the crude product was dissolved in water and resalted with about half the calculated amount of sodium chloride necessary to give a saturated solution. This procedure was fairly efficient, but of course some sodium chloride was still a contaminant. In a second procedure, the crude product was dis- solved in a small amount of anhydrous methanol and filtered to re- move any insoluble material, including sodium chloride. A 26 fifteen-fold volume of ether was then added to reprecipitate the in- dophenol. This procedure was not satisfactory, however, because in many cases decomposition of the indophenol occurred in non-aqueous solutions. No truly satisfactory method for isolating high purity sodium salts of indophenols was devised. In the only published values for the visible spectra of indophenols as their sodium salts, the indophenols were prepared in solution by the reduction of a 4- ‘ aminophenol with ferricyanide to the quinone monoimine, which then coupled with phenol in basic media to form the indophenol salt. In- dophenols prepared in this research have absorption maxima in good agreement with the published values, but their molar absorptivities are 10-15% lower. This difference nay well be accounted for by contamination of the indophenols with sodium chloride. It was observed that when both the quinone-N-chloroimine and phenol were 3,5-disubstituted, no indophenol was produced. Both 3,5- dimethyl and 2,3,5,6-tetramethylquinone-N-chloroimines produced the blue color charactersitic of indophenols when added to alkaline phenol, so steric hinderance, rather than a lack of reactivity, was assumed to inhibit the formation of these symmetric indophenols. Preparation of Other Materials Synthesis of sodiumghypochlorite Sixty-three grams of sodium hydroxide were dissolved in 100 m1 of water, cooled, and poured onto enough crushed ice to give a total weight of 1 kg. Next, tank chlorine diffused through a 27 glass frit was bubbled through the solution until the weight had in- creased by 75 g. The resulting solution was pale yellow and still contained a large amount of ice. Synthesis of monochloramine Twenty grams of ice, 70 m1 of ChloroxR (2 0.7 M_Na0Cl) and 100 m1 of ether were combined in a 500 m1 separatory funnel. After about ten minutes, the contents of the separatory funnel had cooled to 2°C, and 7.5 m1 of concentrated (2 7.4 M) ammonium hydroxide were added. The reaction mixture was gently shaken, after which the aqueous phase was discarded. Then 100 ml of ice cold distilled water were added. After gentle shaking, the aqueous phase was trans- ferred to a brown bottle and refrigerated. Synthesis of sodium aguopentagyanoferrate Solutions containing 20 g of sodium nitroferricyanide in 60 m1 of distilled water, 8 g of sodium hydroxide in 20 ml of distilled water, and 7 g of hydroxylamine hydrochloride in 20 ml distilled water were prepared individually and cooled to -2°C in a salt-ice bath. The sodium hydroxide solution was added to the sodium nitro- prusside solution, and then the hydroxylamine solution was added drop-wise. The reaction mixture turned brownish-green and gas was evolved. At no time did the temperature of the reaction mixture rise above 0°C. After one hour, 300 m1 of ice cold ethanol was added, and the aquoferrate precipitated as a brown resinous mass. The liquid was decanted, and the solid was redissolved in about 60 28 m1 of ice cold distilled water. The aquoferrate was again pre- cipitated by the addition of 200 m1 of ice cold methanol and suction filtered through a medium glass frit. After several more dissolu- tions with distilled water and reprecipitations with methanol, the aquoferrate was filtered, washed with a little ether, and trans- ferred to a vacuum desiccator over concentrated sulfuric acid. The dry product was a gold powder. Discussion Sodium hypochlorite was prepared by the procedure of Raschig (32). The procedure was straightforward but required a rather cum- bersome, large-capacity, double-pan balance to monitor the weight changes in the reaction mixture. When used for qunione-N-chloroimine synthesis, this reagent was prepared just prior to its use and still contained some crushed ice. Monochloramine was prepared by the procedure of Kleinberg, gt,_a1, (21). The concentration of each batch, as determined by thiosulfate titration, varied slightly but was usually about 0.06 M, Due to the instability of this solution, it had to be prepared freshly every few days. Sodium aquopentacyanoferrate was prepared by the procedure of Hofmann (17) as modified by Asperger, gt,_al, (l). The tempera- ture had to be maintained at or below 0°C throughout the synthesis and for the first few recrystallizations to prevent the formation of iron oxides. Due to the resinous nature of the product in prelimi- nary precipitations, decantation, rather than filtration, for the removal of the supernatant liquid was more expedient. V. NON-KINETICS EXPERIMENTS UV Spectrum of Hypochlorous Acid The ultraviolet absorption spectrum of hypochlorous acid as a function of pH is shown in Figure 1. Free hypochlorous acid is characterized by a single band at 233 nm with a molar absorptivity of 134 2 mole-1cm']. Hypochlorite ions display a broad band with maximum absorption at 293 nm and a molar absorptivity (above pH 11) 1cm-l of 358 2 mole" Under conditions of constant pH, solutions of hypochlorite ions obey Beer's Law in the concentration range of 10'5 to 10"3 molar. The ultraviolet absorption spectrum of sodium hypo- chlorite buffered at pH 11.5 is shown in Figure 2. UV Spectrum of Monochloramine Solutions of monochloramine display a broad absorption band between 200 and 300 nm with Amax at 243 nm and a molar absorptivity of 460 z mole'1cm']. Dilute solutions of monochloramine decompose slowly but are relatively stable in the pH range of 7.5 to 11.5. The ultraviolet absorption spectrum of monochloramine buffered at pH 10.0 is shown in Figure 3. 29 30 0-5 F no 0.4 0.3 0.2 O.| 0-0 1 1 l 1 n I 220 240 260 280 300 320 340 Mnm) Figure 1: UV Spectrum of Hypochlorous Acid as a Function of pH. LC) (>13 (3J3 ()3? CXGB (345 (All (3J3 (X23. Ct! ()1) IIIIIIIIIII 220 250 280 310 340 Figure 2: Mnm) UV Spectrum of Hypochlorite Ions 31 LC) ()1? - CRIB- (337 ~ (LES (315 (3J4 - (3J3 - 10.2?- (D.l - ()1) 220 250 280 Figure 3: I l' I I] I 3IO 34o Mnrn) UV Spectrum of Monochloramine 32 Visible Spectra of Pentacyano Iron Complexes The absorption maxima of the pentacyano iron complexes used in this research are listed in Table 1. When monochloramine is added to aqueous solutions of any of these complexes, a rapid shift in the wavelength of maximum absorption to 392 nm occurs. An iden- tical shift occurs when hypochlorous acid is added to any of these complexes. Literature values for Amax and log Emax are shown in parentheses. TABLE 1 Absorption Maxima and Molar Absorptivities of Some Pentacyano Iron Complexes Compound Xmax(nm) log emax NaZEFe(CN)5N0] 398 1.30 (398) (1.31) Na4[Fe(CN)5N02] 398 2.92 (398) (2.99) Na3[Fe(CN)5H20] 415 2.29 (428) (2.40) Visible Spectra of Indophenols The visible absorption spectrum of 2,6,2',6'- tetramethylindophenol as a function of pH is shown in Figure 4 and 33 0.9 P "'° 0.8 '- 07 L 0.6 *- 90 0.5 -' A 0.4- 0.3 - 7.0 0.1 *- 6.0 5.0\. 0-0 In I I I T I I I I I. 460 500 540 580 620 660 700 X (nm) Figure 4: Visible Spectrum of 2,6,2',6'—tetramethylindophenol as a Function of pH. 34 is typical of indophenols. The free indophenol which predominates in acidic solutions is characterized by a broad band between 400 and 500 nm, while the indophenolate ion, which predominates in basic so- lutions, displays a more intense band between 600 and 700 nm. The wavelengths of maximum absorption and the molar absorptivities of indophenols synthesized in this research are listed in Table 2 along with the literature values, which are shown in parentheses. Values of Amax agree quite favorably with the literature, and the approxi- mately 10% lower values for Emax are attributable to sodium chloride impurities, as discussed in Chapter IV. TABLE 2 Absorption Maxima and Molar Absorptivities of Some Indophenols and Their Sodium Salts Indophenol Indophenolate Substituents Amax log emax Amax log emax none 495 3.40 635 3.95 (496) (3.83) (637) (4.50) 2,2'-dimethyl 495 3.64 648 4.15 (495) (3.97) (648) (4.52) 2,6,2',6'-tetramethyl 493 3.50 658 4.14 ~ (494) (3.82) (662) (4.47) 2,2'-dichloro 508 3.04 650 4.41 (no data) (no data) (662) (4.47) 2,6,2',6'-tetrachloro 515 3.15 668 4.02 (no data) (no data) (no data) (no data) 35 Formation of Indophenols in Aqueous Solution When solutions of monochloramine, pentacyanoaquoferrate, and phenol buffered at pH 11.5 were mixed, the blue complex which formed displayed a visible absorption spectrum identical to that of indo- phenol. As the molar ratio of pentacyanoaquoferrate to monochlor- amine approached l:l, the wavelength of maximum absorption remained constant while the absorbance approached a maximum. Molar ratios of pentacyanoaquoferrate to monochloramine exceeding 1:1, however, re- sulted in progressively increasing hypsochromic shifts in the wave- lengths of maximum absorption which approached 700 nm. The same trends were observed when phenol in the reaction mixture was re- placed with methyl- or chloro-substituted phenols. The data from this experiment are presented in Table 3. A similar experiment was performed in which ammonia and a five-fold molar excess of sodium hypochlorite was used in place of the monochloramine solution. Data from this experiment are presented in Table 4. Both data sets are in excellent agreement. These data indicate that only one mole of pentacyanoaquofer- rate is required for each mole of indophenol produced. Furthermore, pentacyanoaquoferrate in excess of this stoichiometric concentration reacts with indophenol or one of its intermediates to form an uni- dentified complex. This complex is not stable, however, and after several days its green color fades into blue, and the visible ab- sorption spectrum again corresponds to that of indophenol. The fate 36 TABLE 3 Effect of Aquopentacyanoferrate Concentration on the End Product of the Berthelot Reaction - -fi— 3.. _ [Fe(CN) H 0] [Fe(CN)5H20]3 [HZNCl] 5 2 max A* at Amax [HZNCl] 1.05 x 10'5 8.52 x 10"5 1.23 x 10" 635 1.076 5.26 x 10"5 8.52 x 10'5 6.17 x 10'1 635 1.680 1.05 x 10‘4 8.52 x 10'5 1.23 641 1.520 5.26 x 10"4 8.52 x 10'5 6.17 688 1.348 1.05 x 10"3 3.52 x 10‘5 1.23 x 101 700 1.426 *1 cm cells TABLE 4 Effect of Aquopentacyanoferrate Concentration on the End Product of the Berthelot Reaction in the Presence of Hypochlorite Ions -_.._.-..—.._. .-.._...~-.—._ “—_. —.—_-._ 7...... - ._____——_.._ _._ m.“ .-——..-—.--—. :— _.—._. -———_—.—.._..— -—. .——__-...—_ --— . .—......—.-.-._._ ._ .. ___ Fe CN H 0 3' A+ at [Fe(CN)5H20]3- [NH3] [-OCIJ E EH2;:]§*] max Amax 1.05 x 10'5 2.47 x 10'4 1.49 x 10'3 4.25 x 10'2 633 1.596 5.26 x 10'5 2.47 x 10"4 1.49 x 10‘3 2.13 x 10" 635 3.266 1.05 x 10'4 2.47 x 10'4 1.49 x 10'3 4.25 x 10"3 635 3.560 5.26 x 10'4 2.47 x 10"4 1.49 x 10'3 2.13 675 2.614 1.05 x 10'3 2.47 x 10'4 1.49 x 10'3 4.25 695 2.926 *Total conversion of ammonia in solution to monochloramine is as- sumed. +l cm cells 37 of pentacyanoaquoferrate in this process was not investigated. It should be noted, however, that when solutions containing synthetic indophenol and a large excess of pentacyanoaquoferrate were mixed, no shift in wavelength was observed. Identity of the "Catalyst" for the Berthelot Reaction Methods of ammonia analysis in which sodium nitropentacyano- ferrate (nitroprusside) or sodium nitritopentacyanoferrate is used to catalyze the Berthelot reaction often produce colors with a greenish tint, rather than the clear blues characteristic of indo- phenols. No clear-cut correlation for the stoichiometric amounts of these compounds necessary to produce the hypsochromic shift observed for sodium aquopentacyanoferrate could be made, but in general, 100 to 150 times as much was required. This led to the hypothesis that sodium nitropentacyanoferrate and sodium nitritopentacyanoferrate were in equilibrium with sodium aquopentacyanoferrate. This hypoth- esis was confirmed by the work of Swinehart and Rock (41), who re- ported rapid conversion of the nitro complex to the nitrito complex in alkaline solutions. After this initial reaction (I), they ob- served that a second equilibrium (II) was established between the [Fe(CN)5N0]2' + 20H" # [Fe(CN)5N02]4' + H20, req = 1.5 x 106 (I) [Fe(CN)5N02]4' + H20 ‘7: [Fe(CN)5H20]3' + N02", Keq = 3 x 10'4 (II) . .ll. [II-I'll"! 38 nitrito and the aquo complexes. These observations compled with the simple 1:1 stoichiometry observed between monochloramine and sodium aquopentacyanoferrate in indophenol formation support the hypothesis that the aquo complex is the actual catalyst for indophenol forma- tion in the Berthelot reaction. VI. KINETICS EXPERIMENTS Formation of Monochloramine The rate of formation of monochloramine as a function of pH were obtained by mixing buffered, equimolar aqueous solutions of am- monium chloride and sodium hypochlorite in the stopped-flow spectro- photometer and recording the change in absorbance with time at either 243 nm or 292 nm, which are the wavelengths of maximum ab- sorption for monochloramine and hypochlorite ions, respectively. At both wavelengths, the reaction was found to be second-order overall and first-order each in hypochlorous acid and ammonia. The max- imum observed rate constant occurred at pH 8.0. The magnitude of the observed rate constants calculated from data collected at 243 nm was approximately twice as large as those calculated from data col- lected at 292 nm for pH 3 10.0. This is because the band for mono- chloramine at 243 nm is overlapped by the hypochlorous acid band at 233 nm, the dichloramine band at 230 nm, and the tailing edge of the hypochlorite ion band, while the hypochlorite ion band at 292 nm is not overlapped by any other species present in the reaction mixture. For this reason, the data collected at 292 nm should give the best estimate for observed rate constants. Data for the formation of monochloramine when the molar ratio of ammonia to hypochlorous acid was 1:5 and 1:10 were also collected at 292 nm over the same pH 39 40 range. Computer plots of data collected at pH 8.0 when solutions of ammonia and hypochlorous acid with molar ratios of 1:1, 1:5, and 1:10 were mixed in the stopped-flow spectrophotometer are presented in Figures 5, 6, and 7, respectively. On the plots, "x" represents experimental points, "0" represents points calculated from the data assuming general second-order kinetics, and "=" represents the over- lap of experimental and calculated points. Each point plotted is the average of ten separate determinations. Values for the observed second-order rate constant agree within 10%. Data from these ex- periments are presented in Figure 8, where the log of the observed second-order rate constant is plotted against pH. As indicated in Figure 8, the observed second-order rate constants for monochloramine formation are highly pH dependent. The fact that the maximum rate occurs at pH 8.0 indicates that the species involved in monochloramine formation are ammonia and hypo- chlorous acid rather than ammonium ions and hypochlorite ions. This is most easily visualized with the aid of a distribution diagram for ammonia, ammonium ions, hypochlorous acid, and hypochlorite ions as a function of pH, such as the one presented in Figure 9. As can be _seen in Figure 9, pH 8.0 corresponds to the maximum concentrations of ammonia and hypochlorous acid relative to each other. If the re- action involved ammonium ions and hypochlorite ions, the maximum ob- served rate constant would occur at pH 9.5. 41 . -o_ x mo.F u “cmemaucw , -o_ x ¢S.N n 268088 “6 8=_8> , -o_ x so.“ u no» we maps» .5: «mm um wocmagomam mw mvxm Pmueggm> .m-o_ x em.m acmemcucp m72 x em.m mipcm_2 pm m=_8> , -o_ Xem.m Hipemp 88 6=_8> .mmv as?“ m_ 85x8 P8p=o~waoz .E 3.0. x me._ u .m-_ooU ecu .2 e-o_ x me._ u . mzzH .e: Nam u x .o.m n :8 .owpmm LwFOZ Fu— cw ku< maogopguoaxz ucm mwcoss< sage mcwEmgopgoocoz we cowpmsgom ”m wgzmwm HF- :uTa UU—wur_u upmfi..t:nn>r\q_?:$.¢ e oeseaaaoooereeoen- _Uiii1411113111IHIIIIJIIIIXIIIIUIIIImIIIIUIIIIUIIIIUIIIIw11IINIIII!JJIIUIIII- 11111 xiiiouililwiIIIUIIII. ”in HLVr n » >Lhnhnr» .r — u,IHIH.I,. H r pH!» 5 p u nun L. n U stun“ u w "MHCX .u. _ 22." 1 .. w ‘ Hun . .— HHH . U "all" u i. » re” 5 . b "C" o. w H" p » nun _ U nu - 1 1 11 -1-..M F Hh.‘ w p .H O u H" o _ up - w w "Cy u » ycx — — cS. » F I - - 1 xcn :11 - .~ » HS » u Tu « » u » . n - m w H" . p u- h n C u h - I I 1.1m1 111 - 111 L. _ u M r us N — .g h v - 1 - p u _ en. M b c h p C u c - i - i 11- x0 - - w v u u _ x w m 2 p2 w 2 - 1 c- - a h re p ... x 0 . u m . .l,1; _ . . -1.-- ".2 p r» 2 on .m1111w11110111101i118 11111 01-11811inviiiaviiIIJiiiiwiiiimiiiixiiiimiiiimiiiiuiiii0111101111m1111m1111~ 42 . -33 x 33.3 - pcmsmaoc3 m -33 x “3.3 1 533333 38 3:333 ,_-o_ x 33.3 n no» we m=_mw .Ec Nmm an mucmagomam we mpxm pmuwugm> . 103 x mm.~ u pamEmLucp "Niop x Fo.~ m1ugmwg um mspm> m io_ x om.o m1u$m3 pm o33m> .mmv wavy m3 mpxm FmpcoNPLo: .E 3-33 x 33.3 n .m-333U .3:3 .E 3-33 x 33.3 - . mzzH .2: N33 u x .o.m u :3 .oppum LmFoz may :3 u30< mzoLo—gooaxz use wwcoes< seem mcpemgo350ocoz yo cowmeLom no mgzmwm 9.0 «110400! mu each m muchnumh Juc1\ntz tattooeeeeeectoeoee- — O cox Cnx 1-1 Cnxx — ccuxx OCIx _- an .-- I 1 ouli - 3 x13 . w xuoc .. _ - xxpc — xxcc u xqu » xllc .Wi xlo : w xnc xfln — xnc . W xnn ; an r u" "no r u 70 . h “H V m- il." _- m — KuC XHC _- W e * "NX m en _ V nux w 3 on 3 * "C W u xmo m w 3" u — "C n _ W b "n ~ "u M V "x. U — Cux u — Onxxu . c _ » O~ -m-U'UWUIUUWUUUUm-UUUUUUUUwUI-UUWUUUUm'UUUVUUUUUUU-IIWUUUUWUUUUVD'5'“-nuUUmUDUDUUUn-UWUI-UUUUUUWUUUUWUUUUn 43 u no» pm «33¢»- m_-o_ x No.N m.3;mmg 33 m=_3> 2 x 33.3 a .m o_ x Nm.3 u 333233333 ”w-op x 33.3 n .5: «am pm mucmagomnm 3 op x omwm mlpmwp no mspm> mwxm Fuowpgm> soupon pm m33m> «3.03 x mo.m .muop x mm.3 u ucmsmgocw .«mv mswp m3 mvxm Fmpcongo: ommzzU .5: Nam u 3 .o.m u =3 .3 - .owpmmgmmo ~m----m---- Mu now""tubululuuunuuuunuuuuuuunuCOOIXxx xxx Cccuuuuflnuunno x I l l S l I '— u-b-o— h-m DID-Inna“ Hunt-Hr unto-Ir- tuba—any our- Int-cu onuo-om‘fu-nut-o—nm I-lO-rt-IO-iu‘ Hun-hut- 1 I V w-1--m----m--'-w-'-lr'--‘V'-’-r-..-'m-'--w-|.II-all-"-u---aim-'"W'---m'--'m--'-w----w-"Im---'m"--m'-. v--1.xm----m---‘W'---m----m----m---.llmu.jul' .111'm---” - FUOH 3:3 .3 -33 x 33.3 n z opu_ 33 3333 3333333333>=.w cm upcoea< sogw mcwsmgopguocoz mo cowumELom .0 watt mg I.-- -'--m-'-- quCInII ”n mgsmwm CJOW c— Whhnmwh Aomz\n12 tn...0¢o¢¢.t¢toot¢¢o I---m---'m...,"m----m--1-M ha Int-«pa ”U" *O-O-IU‘O-‘O-ONHUY nwwmI—Ho-o— mo—amWo—ounnm—u—Ifiwu‘ mumwu IO! 44 log kObS l I l I 7.5 8.0 9.0 I0.0 l LO l2.0 pH Figure 8: Log of observed Second-order Rate Constants for the Formation of MonochIoramine versus pH at 292 nm (.) and 243 nm (I). 45 Figure 9: Distribution Diagrams for Hypochlorous Acid-Hypochlorite Ions and Ammonia-Ammonium Ions as a Function of pH. 46 Decomposition of Monochloramine Monochloramine and hypochlorous acid are mutually unstable, as evidenced by the decrease in absorbance at 243 nm and 292 nm when their dilute aqueous solutions are mixed. The effect of pH on the stability of these species was determined by mixing 1.45 x 10'4 ll monochloramine and 1.49 x 10'3 F hypochlorous acid, buffered at the same pH, in the stopped-flow spectrophotometer. The pH values chosen for these experiments were 8.5, 10.5, and l2.0. All experi- mental data presented are the average of four runs. In the first set of experiments, the change in absorbance at 243 nm with time was recorded. A "blank" in which buffered hypo- chlorous acid was mixed with buffer containing no monochloramine was also performed at this wavelength so that the contribution of hypo- chlorous acid and hypochlorite ions to the absorbance could be sub- tracted. In the second set of experiments, the change in absorbance at 292 nm with time was recorded for identical solutions. No blank was necessary for these experiments since hypochlorite is the only species in solution which absorbs significantly at this wavelength. Neither set of data could be fitted to overall second-order kinetics. Furthermore, the data could not be fitted to either second-order in monochloramine, or second-order in monochloramine and first-order in hypochlorous acid. This indicates that the de- composition of monochloramine is a multi-step process. Some possi- ble pathways for monochloramine are outlined below. Reaction (I) is not expected to occur above pH 7.5 (7). Reactions (II) and (III), 47 HZNCl + HOCl ———> HNCl2 + H20 (1) HZNCl + HOCl ——) HZNOH + 2c1' (II) 'OH - (III) however, were reported by Yagil and Anbar (49) as side-reactions in synthesis of hydrazine when the concentration of monochloramine was less than 1043 H, The ammonia formed in reaction (III) would react with any hypochlorous acid remaining in the solution to form addi- tional monochloramine, which would further complicate the kinetics of the overall decomposition reaction. Plots of absorbance at 243 nm (corrected for absorbance of hypochlorous acid and hypochlorite ions) and 292 nm versus time are presented in Figures l0 and ll, respectively. Both data sets are in excellent agreement. The sharp rise in absorbance in Figure 10 with a maximum occurring at 35, 65, and 305 for pH values of 8.5, 10.5, and l2.0, respectively, are corresponded by sharp decreases in ab- sorbance in Figure l0 at the same time intervals. Initial changes in absorbance indicate the formation of monochloramine from small quantities of ammonia in the monochloramine solution. The steepness of these changes reflects the effect of pH on the rate of monochlor- amine formation reported in the previous section. Furthermore, the data show that monochloramine and hypochlorous acid are relatively stable in the pH range of 10.5 to 12.0. 48 p : i d J—M;_¢ .IO- A - .08- .06- -04 l I l I l 20 so IOO ' I40 ' Ie'o ' 22'o ' 260 300 Time (s) Figure 10: Decomposition of Monochloramine as a Function of Absor- bance at 243 nm and Time at pH Values of 8.5 (I), l0.5 (0), and l2.0 (A). 0.56 ‘- o.54 - — fl A .J 0.52 - 0.50- I I I I I I I T I r I I 7 I T 20 so IOO I40 :60 220 260 zoo Time (s) Figure ll: Decomposition of Monochloramine as a Function of Absor- bance at 292 nm and Time at pH Values of 8.5 (I), l0.5 (0), and l2.0 (A). 49 Mechanism and Kinetics of Indophenol Formation The mechanism and kinetics of indophenol formation from am- monia, hypochlorous acid, sodium aquopentacyanoferrate, and phenol are complex. Experiments in which buffered solutions containing phenol, 2—methylphenol, or 2-chlorophenol plus sodium aquopentacy- anoferrate and monochloramine were mixed in the stopped-flow spec- trophotometer were performed. The change in absorbance with time was recorded at 635 nm and 650 nm, which corresponds to the wave- lengths of maximum absorption for indophenol, 2,2'-dimethyl-, and 2,2'-dichloraoindophenol. Representative data are presented in Figure 12. Brief attempts to fit these data to rate laws were un- successful. Despite this fact, the following mechanism is proposed. It is based on the results of this research, related research of other workers, and a fair measure of intuition and speculation. NH3 + HOCl ——+ HZNCl + H20 (IV) HZNCl + [Fe(CN)5H20]3' 3:2 [Fe(CN)5H2NCl]3' + H20 (V) 91 [Fe(CN)5H2NCl]3' + .014 3-=-*—— [Fe(CN)5-N-.—0H] + 3H20 (VI) Cl C [Fe(CN)5-l!l—.—OH]3' + H20 x—"—‘—— [Fe(CN)5H20]2" + :Il Cl l - H0-.-N: + .-0H + 30H —) 0=”=N--O- + C1- + 3H20 (VIII) 1 ‘ou (VII) III IIIIIIIIIIII III I . I I IIIIll 50 .400 .300 .200 ..|00 .000 Figure l2: / 2!) EH) / J KDC) l h4() l l I ISO 220 260 300 seconds Rate of Formation of 2,2'-dimethylindophenol (A), Indophenol (I), and 2,2'-chloroindophenol (0). 51 Reaction (IV) corresponds to the formation of the first in- termediate in the overall reaction. The rate of this step is gov- erned primarily by the pH of the reaction medium, as shown in the previous section. Reaction (V), which corresponds to the formation of the second intermediate, is supported by experiments discussed in Chapter V. Reactions (VI) and (VII) are speculative, but similar reactions between sodium nitropentacyanoferrate and acetophenone re- ported by Holfe and Swinehart (48) lend some support. Reaction (VIII) corresponds to the formation of indophenol and is supported by the work of Corbett (9) and of Gibbs and Clark (l4). Rates of Formation of Several Indophenols Referring again to Figure 12, it can be seen that the rate of formation of 2,2'-dimethylindophenol is greater than that of in- dophenol and that the rate of formation of 2,2'-dichloroindophenol is less than that of indophenol. These observations are easily rationalized in terms of the inductive effects discussed in Chapter II. Despite the fact that the rate of formation for 2,2'- dichloroindophenol was the slowest, its equilibrium value for ab- sorbance was always the highest. This is probably because its molar absorptivity is the highest of the three. As an aside, 2,6- dimethylphenol coupled much faster than 2-methylphenol, and 2,6- dichlorophenol coupled much slower than 2-chlorophenol. The use of these phenols was hindered, however, due to their low solubility in 52 water. Furthermore, 2,6-dimethylphenol was rapidly chlorinated by hypochlorous acid, which resulted in yellow-orange oils. It was interesting to note that two ortho substituents on a phenol exerted effects greater than would be predicted by simply doubling the ef- fect of one ortho substiuent. The relative rates of coupling for these phenols are as follows: CH3 CH3 Cl CI ‘OH >> ‘OH > .OH > ‘OH >> ‘OH CH3 Cl VII. A NEW AMMONIA METHOD In this chapter, a method for ammonia analysis which evolved from a mechanistic consideration of the Berthelot reaction is pre- sented. The order of reagent addition has been reversed relative to most procedures, and 2-chlorophenol was used in place of phenol. Sodium aquopentacyanoferrate was used to catalyze the reaction. Reagents Ammonia standards (A) 7.44 x 10'1 M; Dissolve 3.9797 g of ammonium chloride (dried for two hours at l20°C) in lOO ml deionized distilled water. (B) 7.44 x 10'4 11: Dilute 1 m1 of A to 1 9. with deionized distilled water. (C) 7.44 x 10'6 _11: Dilute 1 m1 of 13 to 100 m1 with deion- ized distilled water. (D) 7.44 x lO'5 M; Dilute 10 ml of B to lOO ml with deion- ized distilled water. When l ml, 3 ml, and 5 ml of solution C were diluted to 25 -7 ml, the concentrations of ammonia were 2.98 x 10'7 H, 8.93 x 10 fl, and 1.49 x 10'6 fl, respectively. When 1 ml, 3 ml, and 5 ml of solu- tion D were diluted to 25 ml, the final concentrations of ammonia 6 6 5 were 2.98 x 10' M, 8.93 x 10' M, and l.49 x 10' M, respectively. 53 54 Sodiumghypochlorite (E) 7.44 x 10"2 R) M; Dilute l0 ml of 0.744 M_sodium hypo- to 100 ml with deionized distilled water. 2 chlorite (Chlorox (F) l.49 x 10' M; Dilute l0 ml of solution E to 50 ml with deionized distilled water. Sodium aquopentacyanoferrate (e) 4.14. x 10'3 M; Dissolve 0.03 g of Na3[Fe(CN)5H20]'H20 25 ml of deionized distilled water. 4 (H) l.66 x lO' M; Dilute l ml of solution G to 25 ml with deionized distilled water. 2-chlorophenol (I) 1.06 M; Dissolve 2.3 g of 2-chlorophenol in 20 ml of 95% ethanol. (J) 1.05 x 10“ 1_1_: Dilute 5 m1 of solution I to so ml with deionized distilled water. When l ml each of reagents F, H, and J were diluted to 25 ml, their final concentrations were 5.95 x 10'4'M, 6.62 x lo-6 3 M, and 4.24 x lO' M, respectively. The Method Into twelve 25 ml volumetric flasks, which had been rinsed with l0% HCl and deionized distilled water just prior to use, were pipetted the required amount of standards C and D to give ammonia 7 7 11.1.49 x 10'6 n. concentrations of 2.98 x lO- M, 8.93 x l0" 55 —6 6 5 2.98 x 10 M, 8.93 x 10' M, and l.49 x 10' M_when diluted to 25 ml. Next, 15 ml of pH l0.5 buffer was pipetted into each flask, as well as two other 25 ml volumetric flasks containing no ammonia, which served as reagent blanks. Then 1 ml of reagent F was added and the flask was swirled. After one minute, 1 ml each of H and J were added in rapid succession, and the mixture was swirled, diluted to the mark with deionized distilled water, and shaken. This pro- cedure was repeated for each flask. The flasks were allowed to stand for two hours at room temperature after which the absorbance of each solution was measured in 5 cm cells at 650 nm against pH l0.5 buffer. The results are presented in Table 5. The method for ammonia analysis described by Solorzano (39), which is in widespread use in water chemistry laboratories in this country because of its high sensitivity, was used for comparison with the proposed method. Solorzano's procedure used sodium nitro- pentacyanoferrate as the catalyst. Standard solutions were prepared identically to those used for the proposed method except l5 ml of deionized distilled water was added to each flask instead of buffer. After addition of reagents, the solutions were allowed to stand for two hours, after which the absorbance of each solution was measured in 5 cm cells at 640 nm. The results are presented in Table 6. Discussion The calibration curves for the Solorzano and the proposed method are presented in Figure 13. Both obey Beer's Law in the 7 5 ammonia concentration range of 10' M;to 10' M, The slope and 56 TABLE 5 Data Obtained When Ammonia Was Analyzed by the Proposed Method [NH3] A650 AStd - Ab‘k s so (% REL. so) 0.00 .032 - 0.00 .034 - 2.98 x 10'7 .070 0.034 e .002 (5.98) 2.98 x 10‘7 .067 8.93 x 10'7 .126 0.096 s .005 (4.77) 8.93 x 10'7 .128 1.49 x 10'6 .191 0.162 e .006 (3.92) 1.49 x 10"6 .200 2.98 x 10‘6 .382 0.344 s .007 (2.06) 2.98 x 10’6 .372 8.93 x 10'6 1.022 0.086 e .004 (0.36) 8.93 x 10'6 1.017 1.49 x 10'5 1.533 1.494 e .008 (0.52) -5 1.49 x 10 1.522 Data Obtained When Ammonia Was Analyzed 57 TABLE 6 by the Solorzano Method wfi .- m -—.-——-——.-- -“1 [NH3] A640 .Astd - Ablk 3 so (% REL. so) 0.00 .048 - 0.00 .048 - 2.98 x 10'7 .073 0.024 e .002 (9.03) 2.98 x 10"7 .070 8.93 x 10‘7 .139 0.093 s .003 (3.04) 8.93 x 10'7 .143 1.49 x 10‘6 .202 0.144 e .014 (9.82) 1.49 x 10"6 .182 2.98 x 10‘6 .306 0.274 e .023 (8.39) 2.98 x 10'6 .339 8.93 x 10‘6 .885 0.865 e .040 (4.58) 8.93 x 10'6 .941 1.49 x 10‘5 1.326 1.344 e .093 (6.89) 1.49 x 10‘5 1.456 58 I.4 - / / 1.2 - / 1.0 e 0.8 — 0.6 - 0.4 — ()1? - 1 I 1 l 1 I 1 l 1 P 1 I 1 l o 20 4o 60 80 ICC 120 140 7 [NH3] x l0 Figure l3: Calibration Curves for Ammonia as Determined by the Solorzano (0) and Proposed (I) Methods 59 intercept obtained from a linear least squares fit of the data from 4 and -2.43 x 10'3, the Solorzano and proposed methods are 9.69 x 10 and 1.11 x 105 and 2.28 x 10'3, respectively. This corresponds to a 12.7% increase in sensitivity of the proposed method over the Solor- zano method. An examination of Tables 5 and 6, in which the aver- age and standard deviation of absorbance for each set of standards, shows that the precision of the proposed method is as much as ten times better than the Solorzano method, eSpecially at higher con- centration ranges. The precision of the proposed method increases with increasing ammonia concentration, while the precision of the Solorzano method is low and random for all concentrations. It is interesting to note that the color produced in the Solorzano method fades to green or yellow after twelve hours, while the color pro- duced in the proposed method is stable for at least four days. The small increase in sensitivity for the proposed method relative to the Solorzano method may be largely due to the fact that the molar absorptivity of 2,2‘-dichloroindophenol is slightly higher than that of indophenol. The large increase in precision of the proposed method relative to the Solorzano method, however, is at- tributable to the fact that side reactions have been minimized in the proposed method. For example, in the Solorzano method, reagents used are in several thousand-fold excess of their stoichiometric concentrations, and the hypochlorite reagent is added last. Under these conditions, ammonia must compete for hypochlorous acid with phenol and catalyst molecules, which are present in very large 60 excess. In addition, both the first intermediate and the final pro- duct of the Berthelot reaction, monochloramine and indophenol, are unstable in the large excess of sodium hypochlorite recommended by Solorzano. In the proposed method, the hypochlorite reagent is added first and in only five-fold excess of the stoichiometric amount necessary to convert all the ammonia in the highest standard to monochloramine. 7 5 In the concentration range of 10' M_to 10' M, the time re- quired to convert 99.5% of ammonia present in solution to monochlor- amine was calculated as forty to fifty seconds when an observed second-order rate constant of 2 x 103 2 mole-lsec-1 was assumed and the initial concentration of the hypochlorite reagent was 5.95 x 10"4 M, This means that the first intermediate in the Berthelot reaction is formed before the other reagents are added so side reactions are minimized. Although the proposed method offers increased sensitivity and precision, it is important to stress that real samples were not analyzed. Calcium, iron, and magnesium present in fresh- and sea- water would be precipitated at pH 10.5 causing serious interference. The addition of sodium citrate or EDTA to the buffer, however, should alleviate this problem. The proposed method could be easily automated and after minor modification could become the preferred method for trace ammonia analysis. LIST OF REFERENCES \OCDNQU'l-bw 10. 11. 12. 13. 14. 15. 16. 17. 18. LIST OF REFERENCES Asperger, S., Murati, I., and Pavlovic, D., J. Chem. Soc., 730 (1960). Beckwith, P. M., Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1972. Beckwith, P. M., and Crouch, S. R., Anal. Chem. 44, 221 (1972). Berthelot, M., Repertoire de Chemie Applique 1, 284 (1959). Bolleter, N. T., et. al., Anal. Chem. 34, 592 (1961). Borsook, H., J. Biol. Chem. 119, 481 (1935). Chapin, R. M., J. Am. Chem. Soc. 51, 2112 (1929). Chem. Abs. 44, 5213c (1950). Corbett, J. F., J. Chem. Soc., 1502 (1970). Crowther, A. B., and Large, R. 5., Analyst g1, 64 (1956). Czech, F. W., Fuchs, R. S., and Antczak, H. F., Anal. Chem. 33, 705 (1961). "Dictionary of Organic Compounds," 4th ed., Oxford University Press, New York, 1965. Dye, J. L, and Nicely, V. A., J. Chem. Ed. 48, 443 (1971). Gibbs, H. D., Hall, W. L., and Clark, W. M., "Studies in Oxida- tion Reduction: XIII. Preparation of Indophenols Which May Be Used as Oxidation Indicators," U.S. Public Health Reports, Suppl. No. 69 (1929). Harwood, J. W., and Huyser, D. J., Water Res. 4, 501 (1970). Hodgson, W., J. Chem. Soc., 866 (1932). Hofmann, K. A., Ann. 312, l (1900). Horn, 0. B., and Squire, C. R., Clin. Chim. Acta. 14, 185 (1966). 61 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 62 Horn, D. B., and Squire, C. R., Clin. Chim. Acta. 12, 99 (1967). Houben, J., et. al., Ber. 48, 2750 (1909). Kleinberg, J., Tecotyky, M., and Audrieth, L. F., Anal. Chem. 88, 1388 (1954). %uboc?insky, B., and Zalta, J., Bull. Soc. Chem. Biol. 88, 1363 1954 . Mark, H. B., et. al., Anal. Chem. 41, 848 (1969). Metcalf, W. S., J. Chem. Soc., 148 (1942). Mikhailov, G. I., Inst. Pure Chem. Reagents, No. 16, 83 (1937) translated in Chem. Abs. 84, 37078 (1940). Morita, Y., and Ogure, K., Ippon Kagaku Zasshi 84, 816 (1963). Morris, J. C., and Neil, I., "The Formation of Monochloramine and Dichloramine in Water Chlorination," American Chemical Society Meeting, Detroit, Mich., April 17, 1951. Norris, R. K., and Sternhell, S., Aust. J. Chem. 18, 859 (1966). Ohkuma, S., Yakugaku Zasshi 88, 493 (1960). Ohkuma, S., Yakugaku Zasshi 88, 505 (1960). "Organic Synthesis," Collective Vol I, p. 511, Wiley, New York, 1941. Raschig, F. E., Angew. Chem. 88, 2065 (1907). Reardon, J., Foreman, J. A., and Searcy, R. L., Clin. Chim. Acta. 14, 403 (1966). Roskam, R. T., and de Langen, 0., Anal. Chim. Acta. 88, 56 (1964). Russel, J. A., J. Biol. Chem. 1§§, 457 (1944). Scheurer, P. G., and Smith, F., Anal. Chem. 81, 1616 (1944). Searcy, R. L., et. al., Clin. Chim. Acta. 18, 170 (1965). Smith, L. I., and Irwin, W. B., J. Am. Chem. Soc. 88, 1041 (1941). Solorzano, L., L. & 0. 8, 799 (1969). 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 63 Summerford, W. T., and Dalton, D. N., J. Am. Chem. Soc. 88, 1330 (1944). Swinehart, J. H., and Rock, P. A., Inorg. Chem. 8, 573 (1966). Thomas, P., Bull. Soc. Chim. 11, 797 (1912). Van Slyke, D. D., and Hiller, A., J. Biol. Chem. 188, 499 (1933). ' Vaughn, W. R., and Finch, G. K., J. Org. Chem. 81, 1201 (1956). Vemura, P. 1., and Abe, M., Bull. Chem. Soc. Jap. 18, 59 (1937). Weast, Robert C., Ed., "Handbook of Chemistry and Physics," 48th ed., Chemical Rubber Co., Cleveland, 1967. Neil, W., and Morris, J. C., J. Am. Chem. Soc. 11, 1664 (1949). Wolfe, S. K., and Swinehart, J. H., Inorg. Chem. 1, 1855 (1968). Yagil, G., and Anbar, M., J. Am. Chem. Soc. 84, 1790 (1962). Yamaguchi, R., Macheda, T., and Veki, M., Yakugaku Zasshi 88, 1534 (1969). Yamaguchi, R., Suzuki, H., and Hirashima, H., Chem. Pharm. Bull. 18, 1866 (1970).