throbber
3852 Fendler and Fendler
`The Journal of Organic Chemistry
`Hydrolysis of Nitrophenyl and Dinitrophenyl Sulfate Esters1
`E. J. Fendler and J. H. Fendler
`Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15318, and
`Radiation Research Laboratories, Mellon Institute, Camegie-Mellon University, Pittsburgh, Pennsylvania 15818
`Received May 6,1968
`
`The pH-rate profile for the hydrolyses of o-, m-, p-nitrophenyl and 2,4- and 2,5-dinitrophenyl sulfates is char-
`acterized by a plateau in the pH 4-10 region preceded by a more rapid acid-catalyzed reaction and followed by
`feeble base catalysis. A linear free-energy plot for log k¡, (see ref 11) against the p   of the corresponding phenol
`with a slope of —1.2 was obtained. Sodium chloride, sulfate, and fluoride and copper(II) sulfate and silver per-
`chlorate only slightly affect ka for 2,4-dinitrophenyl sulfate, but sodium perchlorate significantly decreases it.
`Acetonitrile decreases k<¡ for the same ester; dioxane in concentrations up to 80% has the same effect, but, when it
`reaches 90%, the rate is increased by a factor of 70; and dimethyl sulfoxide and  , -dimethylformamide increase
`it exponentially. The effects of electrolytes, acetonitrile, and dimethyl sulfoxide on the molar activity coeffici-
`ents of 2,4-dinitrophenyl sulfate and its hydrolysis transition state have been separated by distribution experi-
`ments. Sodium chloride slightly “salts in,” and sodium sulfate and perchlorate, and the dipolar aprotic solvents
`“salt out” 2,4-dinitrophényl sulfate, but all these salts and acetonitrile destabilize the transition state while di-
`methyl sulfoxide has little effect on it.
`Linear free-energy correlations have been obtained between the acid-
`catalyzed rate constants, hp, and the dissociation constants of the leaving groups for all three acids at all con-
`centrations and temperatures investigated. The slopes of these plots are between 0.22 and 0.26 indicating that
`the acid-catalyzed hydrolysis of these sulfate esters is relatively insensitive to the electron-withdrawing power of
`the leaving group. The catalytic effectiveness of the acids is HjSCh > HClCh > HC1. Plots of log     + Ho
`against Ho + Ch + are linear, but their slopes are different for the different acids. This behavior is rationalized by
`assuming that the acids, in addition to their proton-donating power, exert specific electrolyte effects on both the
`initial and the transition states of these hydrolyses.
`Recently, considerable interest has been shown in
`mechanistic studies concerning the hydrolysis of sulfate
`esters2-6 owing to their biochemical importance.6 The
`spontaneous2 perchloric acid,3 base,2 and amine2 cat-
`alyzed hydrolyses of p-nitrophenyl sulfate and car-
`boxyl group participation in salicyl sulfate hydrolysis4
`have been reported. Lack of sufficient data has frus-
`trated attempts to compare the mechanisms of aryl
`sulfate hydrolyses with those of the much more widely
`It has been sug-
`investigated monoaryl phosphates.7
`gested that a study of the hydrolysis of sulfate esters
`containing good leaving groups might substantiate the
`proposed unimolecular mechanism.2 Although 2,4-
`and 2,5-dinitrophenyl sulfates have such good leaving
`groups, their preparation and consequently kinetic in-
`vestigations of their hydrolysis have so far eluded the
`attention of chemists and biochemists alike. We,
`therefore, have undertaken a systematic investigation
`of the mechanisms of hydrolysis of both mono- and di-
`nitrophenyl sulfates and report the results of a study of
`their neutral and acid-catalyzed hydrolyses.
`
`hr at 35-40° and for approximately 14 hr at room temperature.
`After the mixture was cooled to 0°, it was poured rapidly with
`stirring into cold 4   KOH (50 ml). The precipitate was
`twice with cold 95%
`filtered immediately, washed at
`least
`ethanol, and dried in vacuo.
`The potassium salts of o- and ro-
`recrystallized from 90% ethanol.
`nitrophenyl sulfates were
`The labile dinitrophenyl sulfates were purified by the addition
`of dry acetonitrile to the solid, centrifugation, and rotary evapora-
`tion of the solution followed by the same procedure but using
`a minumum amount of cold aqueous ethanol. Rapid rotary
`evaporation of the solvent at 10-15° and drying in vacuo gave
`shiny white crystals. The purity of all the sulfates was estab-
`lished to be not less than 99% by complete hydrolysis of known
`quantities in aqueous buffer solutions at pH 10.00, assaying the
`amount of phenoxide ion liberated, and by their ir and nmr spec-
`tra.
`Deionized distilled water was used to make up the buffer
`solution and the standard acid and alkali solutions. The pH of
`the buffer solutions was adjusted by addition of hydrochloric
`acid or sodium hydroxide at 25.00° using an Orion Model 801 pH
`meter. The concentrations of the acid solutions were determined
`by titration with standard 0.10 or 1.00 M NaOH (BDH) using
`Lacmoid as the indicator. The alkali solutions were standardized
`by titration with standard 1.00 M HC1 using Lacmoid as the
`indicator.
`1,4-Dioxane was purified by the method of Fieser*
`and stored in polyethylene bottles in a refrigerator. Spectro-
`grade dimethyl sulfoxide,  , -dimethylformamide, and aceto-
`nitrile were
`used without further purification and stored over
`molecular sieve. The aqueous organic solutions were made up
`by using the appropriate volumes measured at room temperature.
`Kinetics.—The hydrolysis was followed spectrophotometrically
`by determining either the phenoxide ion or
`the phenol concentra-
`tion. The phenoxide ion formation was
`followed in the higher
`pH region and in alkaline solutions. At lower pH values and in
`taken of the fact that the extinc-
`the acid region advantage was
`the nitro- and dinitro-substituted phenol
`tion coefficients of
`sulfates are very much smaller than those of the corresponding
`in strong acid solutions, the formation
`phenols. For a few runs
`followed by adjusting the pH of the in-
`of phenoxide ion was
`dividual samples to 9.0 by the addition of buffer solutions. The
`agreed within 3% with those
`rate constants for
`these runs
`formation.
`The
`obtained by direct measurement of phenol
`following wavelengths (  µ) were used-for following the phenoxide
`ion and phenol formation, respectively: o-nitrophenyl sulfate,
`410, 290; ro-nitrophenyl sulfate, 380, 350; p-nitrophenyl sulfate,
`403, 320; 2,4-dinitrophenyl sulfate, 360, 320, and 2,5-dinitro-
`phenyl sulfate, 440, 270. The Beer-Lambert law was obeyed over
`
`(9) L. F. Fieser, “Experiments in Organic Chemistry,” 3rd ed., D. C.
`Heath and Co., Boston, Mass., 1957, p 284.
`
`Experimental Section
`Materials.—p-Nitrophenyl sulfate potassium salt (Eastman)
`was recrystallized from 90% ethanol.
`o-Nitro-, m-nitro-, 2,4-dinitro-, and 2,5-dinitrophenyl sulfates
`were prepared by a modified procedure of Burkhardt and Wood.8
`Chlorosulfonic acid (35 mmol, 2.32 ml) was added dropwise with
`stirring to a solution of dimethylaniline (12.5 mmol, 11.73 ml)
`in carbon disulfide (12.5 ml) at —16 to —21°. The mixture was
`then warmed to 35-40°;
`the appropriate dry phenol (25 mmol)
`was added with stirring; and the mixture was stirred for 1-1.5
`(a) For a preliminary report, see E. J. Fendler and J. H. Fendler,
`(1)
`(b) Supported in part by the U. S. Atomic
`Chem. Commun., 1261 (1967).
`Energy Commission.
`(2) S. J. Benkovic and P. A. Benkovic, J. Amer. Chem. Soc., 88, 5504
`(1966).
`(3) J. L. Kice and J. M. Anderson, ibid,, 88, 5242 (1966).
`(4) S. J. Benkovic, ibid,, 88, 5511 (1966).
`(5) R. W. Hay and J. A. G. Edmonds, Chem. Commun., 967 (1967).
`(6) K. S. Dodgson, B. Spencer, and K. Williams, Biochem. J., 64, 216
` . Bostrón, K. Bernstein, and M. W. Whitehouse, Biochem. Pharma-
`(1956);
`col., IS, 413 (1964).
`(7) J. R. Cox, Jr., and O. B. Ramsay, Chem. Rev., 64, 317 (1964).
`(8) G.  . Burkhardt and H. Wood, J. Chem. Soc,, 141 (1929).
`
`MAIA Exhibit 1034
`MAIA V. BRACCO
`IPR PETITION
`
`

`

`Vol. 33, No. 10, October 1968
`
`Hydrolysis of Nitrophenyl Sulfate Esters 3853
`
`the range of concentrations employed for all the sulfates. Ki-
`netic runs at 14.90, 25.00, and 45.00° were
`followed directly
`in the thermostated cell compartment of a Beckman DU-2
`carried
`spectrophotometer. Those at 75.00 and 100.00° were
`out in sealed tubes which were placed in a thermostat, individually
`withdrawn at the required times, quickly cooled, and analyzed
`For
`the appropriate wavelength.
`spectrophotometrically at
`the hydrolyses in sodium hydroxide solutions, alkali resistant
`glass (Corning No. 7280) was used. The temperature in the
`thermostats and inside the cell compartment was measured by an
`NBS thermometer and maintained to be ±0.02°. Examples of
`are shown in Figure 1.
`typical kinetic runs
`Partitioning Experiments.—The measurement of the distribu-
`tion of 2,4-dinitrophenyl sulfate between water and an immiscible
`temperature and in the presence
`organic solvent, at a constant
`and absence of salts or of other media, gives the effect of the
`electrolytes or of the media on the molar activity coefficient of
`the sulfate ester.10 Water and cyclohexane, water and hexane,
`cyclohexane and dimethyl sulfoxide, and hexane and acetonitrile
`the salts are not extracted into
`are almost completely immiscible;
`the organic layer; and the density differences of these solvents
`readily allows them to be separated quickly and cleanly after
`shaking. The hydrolysis of 2,4-dinitrophenyl sulfate at 25.00°
`was negligible compared with the distribution time. Even in 75%
`dimethyl sulfoxide, representing the fastest hydrolysis rate, the
`half-life for the hydrolysis of 2,4-dinitrophenyl sulfate is ap-
`proximately 30 min and generally the distribution was carried
`in less than 2 min. Typically a known concentration of 2,4-
`out
`dinitrophenyl sulfate was made up by weight in a buffer or elec-
`trolyte solution (25.00 ml), and the concentration of this solution
`redetermined by measuring the absorption at 320   µ of a
`was
`suitable dilution using the appropriate blank solution. These
`two independently determined concentrations usually agreed
`within ±3%. A 10.00-ml portion of
`the aqueous 2,4-dinitro-
`phenyl sulfate was vigorously mixed with 10.00 ml of cyclo-
`hexane or the same volume of hexane in a thermostated separatory
`funnel at 25.00°. After the two layers had separated, the con-
`centration of 2,4-dinitrophenyl sulfate was measured spectro-
`photometrically in each layer at 320   µ using the appropriate
`blank solution. The distribution coefficient,
`is defined as
`r,
`in hexane/[2,4-
`in cyclohexane or
`[2,4-dinitrophenyl sulfate]
`in aqueous media. The sub-
`in water or
`dinitrophenyl sulfate]
`script o refers to water and s to salt solutions or other media.
`The ratio re/r„ is equal to the ratio /,//„, where /s is the molar
`activity coefficient of 2,4-dinitrophenyl sulfate in the presence
`of electrolytes or other media, and /„ is that
`in water.
`Six
`independent values for r0 were determined, and they agreed
`within ±5%.
`Furthermore, 2,4-dinitrophenyl sulfate was
`distributed between 3.75 Af NaCKX and cyclohexane, and in a
`separate experiment hexane was used instead of cyclohexane.
`Excellent agreement was obtained between the values found in
`these two experiments (Table I).
`
`Table I
`Activity Coefficients of 2,4-Dinitrophenyl Sulfate
`Its Hydrolysis Transition State in the Presence of
`and
`Electrolytes, Acetonitrile, and Dimethyl Sulfoxide0
`Medium
`/a//o
`1.25 Af NaCKX
`1.38
`1.92s
`2.50 Af NaCKX
`1.19
`2.75s
`3.75 Af NaCKX
`1.37
`4.25s
`3.75 Af NaCKXc
`1.38
`4.24s
`0.75 Af NaCl
`0.94
`1.05s
`1.50 Af NaCl
`0.91
`1.00s
`2.25 Af NaCl
`0.82
`0.97s
`0.75 Af NaeSO,
`1.73
`1.70s
`23.40
`22.40
`25.00% CH3CN
`214.00
`147.00
`50.00% CHaCN
`200.00
`372.00
`75.00% CHaCN
`25.00% DMSO
`0.80
`1.85
`50.00% DMSO
`6.23
`1.30
`70.00% DMSO
`0.87
`8.67
`° Using cyclohexane unless otherwise stated; correlated with
`rate constants at 25.00° unless otherwise stated.
`6 Rate con-
`stants measured at 45.00°.
`0 Using hexane.
`(10) F. A. Long and W. F. McDevit, Chem. Rev.. 61, 119 (1952).
`
`Figure 1.—Plot of 1 + log (OD« — ODt) against time for the
`neutral hydrolysis of sulfate esters. A =
`2,4-dinitrophenyl
`sulfate; pH 5.90, 75.00°. B = o-nitrophenyl sulfate; pH 8.00,
`time scale in 102. C = 2,5-dinitrophenyl sulfate; pH
`100.00°;
`5.90, 100.00°.
`
`Figure 2.—Variation of fa with pH for o-nitrophenyl sulfate at
`100.00°.
`
`Figure 3.—Plot of 8 + log fa at 100.00° against pK„ of the cor-
`responding phenol.
`Results
`for the neutral11
`The first-order rate constants, k0,
`hydrolysis of nitrophenyl and dinitrophenyl sulfates
`The pH-rate profile for o-nitro-
`are given in Table II.
`phenyl sulfate at 100.00° is in Figure 2. Figure 3
`the neutral hydrolysis refers to the hydrolysis of the
`In this paper,
`(11)
`sulfate eeters, in the plateau region of the pH-rate profile, i.e., from pH 4 to
`12, and is symbolized by fco.
`
`

`

`3854 Fendler and Fendler
`
`The Journal of Organic Chemistry
`
`Table II
`Hydrolysis of Nitrophenyl and Dinitrophenyl Sulfates in the pH Region
`10*ti, sec-1"-----
`
`,
`
`----------
`
`—
`
`...
`
`..
`
`—
`
`o-Nitrophenyl sulfate
`100.00°
`75.00°
`1307.0
`306.0
`178.0
`120.7
`40.2
`6.78
`
`pH
`Buffer®
`0.10 M HC1
`1.00
`0.01    ,  ,
`2.50
`0.01 M KH2P04
`2.70
`0.01 M KH2P04
`3.00
`0.01 mkh2po4
`3.50
`0.01 M KHC8H404
`4.00
`0.01 M KHCsHA
`4.32
`5.90
`6.80
`0.01 MKHC,H404
`0.01 Af KH2P04
`7.10
`6.00
`0.01 Af KH2P04
`6.85
`7.00
`8.00
`0.01 Af Na2B407
`7.17
`0.01 Af Na2B407
`9.00
`6.80
`0.01 Af Na2B407
`7.45
`10.00
`0.01 Af K2HP04
`7.20
`11.00
`0.10 Af NaOH
`8.40
`13.00
`1.0 Af NaOH
`68.10
`14.00
`“ The pH of buffers were adjusted at 25.00°.
`
`0.659
`
`0.642
`
`0.644
`
`pKa of the phenol.
`Figure 4.—Plot of 6 + log k$ at 1.00 Af acid against pXa of
`the corresponding phenol at 25.00°: O, HC10«; B, HC1;
` ,
`H2S04.
`
`shows the linear free-energy relationship between the
`dissociation constant of
`the leaving group (nitro or
`dinitrophenol) and the rate constants for the neutral
`hydrolysis. The Arrhenius parameters for the neutral
`hydrolysis are collected in Table III.
`The effects of
`the neutral hydrolysis of 2,4-dinitrophenyl
`salts on
`sulfate at 45.00° are summarized in Table IV, and
`Table V gives the solvent effects on the hydrolysis of
`the same compound. The effects of electrolytes, ace-
`tonitrile, and dimethyl sulfoxide on the molar activity
`coefficient of 2,4-dinitrophenyl sulfate and on that of
`its hydrolysis transition state are collected in Table I.
`The first-order rate constants, k+, for the acid-cat-
`alyzed hydrolysis of nitrophenyl and dinitrophenyl
`sulfates are given in Table VI. Good linear free-energy
`correlations have been obtained between k+ and the
`dissociation constants of the leaving groups for all three
`the concentrations and temperatures in-
`acids at all
`vestigated. A typical plot for a 1.00 M acid at 25.00°
`is given in Figure 4. The slopes of these plots are
`0.22-0.26. Plots of log fc* against Ho12 generally gave
`taken from M. A. Paul and F. A. Long, Chem.
`(12) Values for Ho were
`Rev., 67, 1 (1957), and when available from more
`recently determined values
`such as those which have been reported by M. J. Jorgesen and D. R. Hartter,
`J. Amer. Chem. Soc., 86, 878 (1963), and K. Yates and H. Wai, ibid., 86, 5408
`(1964).
`
`m-Nitrophenyl
`sulfate
`100.00°
`948.0
`
`2,4-Dinitrophenyl sulfate
`45.00°
`75.00°
`326.0
`269.0
`
`2,5-Dinitrophenyl sulfate
`100.00°
`75.00’
`3200.0
`1680.0
`
`72.5
`
`0.31
`0.32
`0.32
`0.31
`0.32
`
`3700.0
`3680.0
`3720.0
`
`295.0
`
`279.0
`278.0
`272.0
`273.0
`
`150.0
`148.0
`149.00
`
`1220.0
`1250.0
`
`1180.0
`1180.0
`
`1160.0
`
`1510.0
`
`Table III
`Arrhenius Parameters for
`the Neutral Hydrolysis
`of Nitrophenyl and Dinitrophenyl Sulfates
`AS*, eu“
`Sulfate
`E, kcal/mol
`-17.4
`24.7
`o-Nitrophenyl
`18.5*
`p-Nitrophenyl
`—18.  8
`-18.0
`18.8
`2,4-Dinitrophenyl
`-17.4
`19.4
`2,5-Dinitrophenyl
`Calculated at 75.00°.
`6 See ref 2.
`
`NasSO,
`CuS04
`AgC104
`NaF
`
`NaCl
`
`NaC104
`
`Table IV
`Salt Effects on the Neutral Hydrolysis of
`2,4-Dinitrophenyl Sulfate at 45.00°
`Concn of salt, M
`106  ·, sec-1
`0.00
`27.9
`1.00
`25.7
`2.00
`25.0
`24.8
`3.00
`20.3
`1.00
`2.00
`15.2
`3.00
`11.00
`4.00
`8.19
`5.00
`5.78
`1.00
`28.0
`0.1
`33.0
`0.1
`35.0
`0.5
`29.5
`1.0
`32.2
`lines whose slopes are given in Table
`good straight
`VII.
`The plot for o-nitrophenyl sulfate is illustrated
`in Figure 5. The scatter between the different acids
`might be due to the different electrolyte effects of the
`acids on the sulfate esters or on their hydrolysis trán-
`sition states.
`'¡To test this point we replotted oür ex-
`perimental datá as suggested by Bunnett18 and later
`modified by Bunnett and Olsen.14 Plots of log k^ + Ho
`against —log oH,o are curved and are different for the
`different acids; some of them show a minimum. Fig-
`ure 6 shows these plots for o-nitrophenyl sulfate.
`In
`their modified treatment Bunnett and Olsen suggest
`that the slope,  , of plots of log
`+ Ho against H0 +
`
`(13) J. F. Bunnett, ibid., 88, 4956 (1961), and accompanying papers.
`(14) J. F. Bunnett and F. P. Olsen, Can. J. Chem., 44, 1917 (1966).
`
`

`

`Vol. S3, No. 10, October 1968
`
`Hydrolysis of Nitrophenyl Sulfate Esters 3855
`
`Table V
`Neutral Hydrolysis or 2,4-Dinitrophbnyl Sulfate in Aqueous Organic Solvents
`Volume %
`E,
`organic solvent
`kcal/mol
`18.8
`0
`18.6
`25
`19.2
`50
`18.4
`75
`20.0
`40
`22.3
`60
`24.9
`80
`25.9“
`90
`18.0
`25
`20.0
`50
`19.4
`75
`19.5
`85
`22.2
`25
`22.1
`50
`= 76 sec-1 at 14.90°.
`
`10*  , sec-1, at 45.00°
`27.9
`27.4
`21.0
`13.1
`50.0
`49.0
`42.0
`
`62.1
`142.0
`302.0
`1130.0
`90.8
`183.0
`
`10*i^, sec-1, at 25.00°
`4.00
`3.85
`2.74
`1.86
`5.94
`4.60
`3.50
`260.00
`9.20
`18.00
`38.7
`143.7
`8.62
`17.6
`
`Medium
`CHjCN-HjO
`
`Dioxane-HsO
`
`DMSO-HjO
`
`DMF-HjO
`
`1 Calculated from 1
`
`AS*, eu, at 25.00°
`-18.0
`-18.5
`-16.0
`-15.5
`-10.0
`-5.2
`0.1
`14.01
`-18.0
`-12.0
`-10.5
`-6.5
`-3.7
`-3.2
`
`Table VI
`Acid-Catalyzed Hydrolysis of Nitrophenyl and Dinitrophenyl Sulfates1
`
`HC1
`
`HCIO,
`
`HiSO,
`
`Concn of acid, M
`1.03
`1.03»
`2.06
`3.06
`4.12
`5.15
`5.15»
`6.18
`7.12
`1.00
`1.00»
`2.00
`3.00
`4.00
`5.00
`5.00»
`6.00
`7.08
`1.00
`1.00»
`2.00
`3.00
`4.00
`5.00
`5.00»
`6.00
`1 At 25.00°, unless specified otherwise.
`
`&copy;-Nitrophenyl
`sulfate
`1.58
`17.8
`4.44
`11.2
`24.5
`48.6
`
`126.5
`258.0
`1.70
`22.2
`4.16
`9.92
`24.0
`60.1
`
`191.0
`759.0
`1.76
`20.1
`4.33
`11.2
`33.6
`92.4
`
`252.0
`» At 45.00°.
`
`m-Nitrophenyl
`sulfate
`0.91
`12.0
`3.26
`6.47
`15.0
`34.3
`387.0
`79.5
`201.0
`0.958
`13.0
`3.06
`5.67
`12.8
`28.0
`297.0
`78.7
`
`0.98
`11.6
`2.94
`7.13
`20.3
`54.1
`456.0
`151.0
`
`p-Nitrophenyl
`sulfate
`2.35
`43.5
`6.30
`18.0
`39.4
`85.5
`
`2,4-Dinitrophenyl
`sulfate
`13.3
`135.0
`29.7
`64.7
`109.0
`276.0
`
`2,5-Dinitrophenyl
`sulfate
`5.09
`45.0
`9.08
`23.3
`52.9
`98.7
`
`184.0
`
`2.35
`45.6
`7.25
`15.2
`35.7
`68.2
`
`228.0
`
`2.36
`44.2
`9.58
`18.5
`51.7
`138.0
`
`651.0
`
`12.7
`126.0
`27.7
`63.6
`142.0
`300.0
`
`862.0
`
`13.2
`151.0
`44.0
`
`225.0
`
`1226.0
`
`293.0
`
`5.17
`53.6
`11.5
`24.5
`52.1
`105.0
`
`307.0
`
`5.27
`49.3
`10.4
`23.4
`69.0
`182.0
`
`743.0
`
`Table VII
`the Acid-Catalyzed Hydrolysis of Nitrophenyl
`  Values1 and Slopes of Log k^ vs. —Ho Plots for
`Dinitrophenyl Sulfates
`p-Nitrophenyl
`m-Nitrophenyl
`o-Nitrophenyl
`sulphate
`sulphate
`sulphate
`-0,05
`-0.07
`-0.03
`HC1
`-0.39
`-0.20
`-0.39
`HCIO,
`-0.20
`-0.15
`-0.20
`,H,SO,
`0.95
`0.94
`0.96
`HC1
`Log k^ vs. —Ho slope
`0.86
`0.85
`0.86
`HCIO,
`at 25.00°
`0.83
`0.85
`0.86
`H,SO,
`log Ch+ + Ho at 25.00° in the 1.00-6.00 M acid range.
`slope of plots of log k+ + Ho vs.
`

`
`1  
`
`=
`
`and
`
`2,5-Dinitro-
`pbenyl sulphate
`-0.10
`-0.38
`-0.14
`1.10
`0.73
`0.73
`
`2,4-Dinitro-
`phenyl sulphate
`-0.10
`-0.46
`-0.32
`0.98
`0.85
`0.86
`
`

`

`3856 Fendler and Fendler
`
`The Journal of Organic Chemistry
`
`-Ha.
`Figure 5.—Plot of 5 + log fa against —H<¡ for o-nitrophenyl
`sulfate at 25.00°:
` , HCIO,; 0, HC1;
` , H2SO,.
`
`Figure 7.—Plot of 6 + log fa + Ho against —
`for o-nitrophenyl sulfate at 25.00°: O, HCIO,;
`H2SO«.
`
`(log Ch+ + Ho)
` ,
`0, HC1;
`
`coefficients of 2,4-dinitrophenyl sulfate to be accurate
`within ± 10%, although the errors
`are smaller for the
`electrolytes. Errors for the transition-state activity
`coefficient ratios are inevitably compounded, but even
`for acetonitrile, representing the most complex system,
`the results are correct within ±25%.
`
`Discussion
`Neutral Hydrolysis.—The pH-rate profile for the
`hydrolysis of aryl sulfate esters involves a plateau, pH
`4-12, preceded by a more rapid acid-catalyzed reaction
`and followed by an exponential curve
`(see Figure 2)
`due to the incursion of base catalysis.15 Superficially
`this hydrolytic behavior is very similar to that which
`has been observed for the hydrolysis of alkyl,7 aryl,15·17
`and acyl18·19 phosphate dianions.
`The present results on the hydrolysis of nitrophenyl
`and dinitrophenyl sulfate anions will be discussed with
`respect to the hypotheses that (a) the rate-determining
`step involves unimolecular sulfur-oxygen bond fission
`with the elimination of S03,
`the hydrolysis pro-
`(b)
`ceeds by bimolecular nucleophilic attack of water on
`sulfur, and (e) water molecules participate in the tran-
`sition state by ion-dipole or dipole-dipole interactions
`or by hydrogen bonding to such an extent that the
`molecularity of the reaction is a matter of its academic
`definition.
`The sensitivity of the rate of hydrolysis to changes
`in substituents in the leaving group (p) has been found
`to be a useful mechanistic probe in studies of phosphate
`ester hydrolysis.16·17·19 We have found that a plot of
`log fa for the sulfate anion hydrolysis against the p  
`is linear with a slope of
`of the corresponding phenol
`— 1.2 (Figure 3). Such a relatively large substituent
`is consistent with a unimolecular mechanism,
`effect
`since in this case the only requirement for hydrolysis is
`sulfur-oxygen bond breaking. However, for the hy-
`drolysis of phosphate monoanions, which exhibits
`many characteristics of a unimolecular reaction but
`(15) The powerful base catalysis at high pH values might involve a change
`from a unimolecular to a bimolecular mechanism involving attack by hy-
`droxide ion on both carbon and sulfur.2 In addition, these processes could
`be complicated by specific electrolyte effects of the alkali metal hydroxide.
`Clearly our data on the base-catalyzed hydrolysis are insufficient to elucidate
`the base-catalyzed mechanism in greater detail.
`(16) C. A. Bunton, E. J. Fendler, and J. H. Fendler, J. Amer. Chem. Soc.,
`89, 1221 (1967).
`(17) A. J. Kirby and A. G. Varvoglis, ibid., 89, 415 (1967).
`(18) G. DiSabato and W. P. Jencks, ibid., 88, 4400 (1961).
`(19) A. J. Kirby and W. P. Jencks, ibid., 87, 3209 (1965).
`
`Figure 6.—Plot of 6 + log fa + Ho against log omo for o-nitro-
`phenyl sulfate at 25.00°:
` , HCIO,; 0, HC1;
` , H2SO<.
`
`log Ch+ indicates the effect of activity coefficient
`ratios on
`the reaction rate.14 Figure 7 shows such
`plots for o-nitrophenyl sulfate. Corresponding plots
`for the other sulfate esters are similar. The slopes of
`these plots are given in Table VII.
`The energies and
`entropies of activation are collected in Table VIII.
`
`Table VIII
`Energies and Entropies of Activation for
`Acid-Catalyzed Hydrolysis of Nitrophenyl
`Dinitrophenyl Sulfates
`
`Sulfate
`o-Nitrophenyl
`
`m-Nitrophenyl
`
`p-Nitrophenyl
`
`2,4-Dinitrophenyl
`
`2,5-Dinitrophenyl
`
`Solvent
`1.03 Af HC1
`1.00 M HCIO,
`1.00 M H2SO,
`1.03 MHC1
`5.15 Af HC1
`1.00M HCIO,
`5.00 M HCIO,
`I.OOMHüSO,
`5.00 M H2SO,
`1.03 MHC1
`LOOM HCIO,
`1.00 M H2SO,
`1.03 MHC1
`LOOM HCIO,
`LOOMHüSO,
`1.03 MHC1
`LOOM HCIO,
`1.00 M H2SO,
`
`B,
`koal/mol
`22.9
`23.9
`21.6
`27.8
`22.4
`25.2
`22.3
`23.4
`20.1
`22.6
`22.2
`23.1
`21.8
`21.6
`22.9
`21.0
`21.2
`21.9
`
`the
`and
`
`AS  , eu
`at 25.00°
`-5.7
`-5.5
`-9.0
`1.0
`-0.4
`1.6
`-1.4
`-1.0
`-7.0
`-2.0
`-6.0
`-1.5
`-1.5
`-5.0
`-1.3
`-8.0
`-6.0
`-7.0
`
`Errors, standard deviations, in individual rate con-
`those in E are
`stants are not greater than ±4%;
`approximately ±0.7 kcal mol-1; and those in AS*
`are ±2 eu. We consider values for the molar activity
`
`

`

`Vol. 33, No. 10, October 1968
`
`the
`involves a proton transfer to the leaving group,
`In this case,
`slope of a corresponding plot is —0.32.20
`a small substituent effect is reasonable since leaving
`groups which weaken the phosphorus-oxygen bond de-
`the basicity of the ester oxygen atom and vice
`crease
`the slope for
`In addition,
`the magnitude of
`versa.
`sulfate anions is considerably greater than seems prob-
`able for a mechanism involving bimolecular displace-
`the bimolecular
`ment by a water molecule.
`Indeed,
`reaction between hydroxide ion and a series of fully
`substrated phosphates with different leaving groups dis-
`plays little sensitivity to the nature of
`the leaving
`group; i.e., a plot of log k against the pK& of the leaving
`group has a slope of —0.43.21 However, the effect of
`substituents in the leaving group on the rate of hy-
`dolysis of nitro and dinitrophenyl sulfates is quite
`similar to that found for the hydrolysis of the dianions
`of substituted acetyl, benzoyl, and aryl phosphates for
`which the corresponding slope is —1.2.16-19 Con-
`siderable evidence supports the conclusion that
`the
`hydrolysis of these phosphates proceeds by unimolecu-
`lar phosphorus-oxygen bond fission with elimination of
`the metaphosphate ion.7 Thus, the available evidence
`for substituent effects suggests that the hydrolysis of
`nitro and dinitrophenyl sulfates proceeds by a uni-
`molecular mechanism, but it does not exclude the al-
`ternative formulation mentioned in hypothesis c.
`The presence of sodium chloride (up to 3.00 M),
`sodium sulfate (1.00 M), and sodium fluoride (1.00 M)
`only slightly affects the rate constant for the neutral
`hydrolysis of 2,4-dinitrophenyl sulfate, but the addition
`of sodium perchlorate significantly decreases it (Table
`I\r). The addition of sodium perchlorate, chloride,
`and sulfate was found to increase the rate of hydrolysis
`of the dianion of 2,4-dinitrophenyl phosphate, but this
`the cation
`effect was more
`specifically dependent on
`than on
`the anion.16 The available data concerning
`the effects of 0.8 M potassium iodide, 0.8 M sodium
`fluoride, and 0.25 M sodium sulfite on the hydrolysis of
`p-nitrophenyl sulfate at 75.00°2 has been interpreted as
`evidence supporting a unimolecular mechanism.22
`However, an intercomparison of these salt effects between
`the various sulfate and phosphate esters is somewhat mean-
`ingless since they can only be described in terms of free-energy
`changes of the initial state relative to the transition state.
`Such changes are best described by the Br0nsted-Bjer-
`rum rate equation which relates the rate constant in the
`presence of electrolytes, k, to the rate constant in their
`absence, k0, and to the ratio of activity coefficients of
`the initial and transition states, fjf*, which follows.
`k = kof./f*
`(1)
`A meaningful interpretation of electrolyte and medium
`reaction rate must necessarily involve a
`effects on
`terms in eq 1
`separation of
`the activity coefficient
`rather than a gross comparison of the observed rate
`constants for seemingly similar reactions. We have
`attempted to do this by partitioning 2,4-dinitrophenyl
`
`(20) C. A. Bunton, E. J. Fendler, E. Humeres, and K-U Yang, J. Org.
`Chem., 32, 2806 (1967).
`(21) D. F. Heath,
`‘Organophosphorus Poisons,” Pergamon Press, New
`York, N. Y., 1961, p 79.
`(22) These salt-effect studies were
`carried out, however, at a constant
`ionic strength of one using potassium chloride which seriously complicates any
`interpretation of
`the effect of 0.8 M added electrolyte on
`the rate of hy-
`drolysis.
`
`Hydrolysis of Nitrophenyl Sulfate Esters
`
`3857
`
`sulfate between cyclohexane or hexane and the aqueous
`solution (see Experimental Section). Combining rjr0
`fjfa with eq 1 one obtains
`
`=
`
`=
`
`/,*//„*
`(2)
`k0rB/kBr„
`which allows separation of the activity coefficient ratios
`for both the initial and transition states. Sodium
`chloride slightly “salts in” and sodium perchlorate and
`sulfate “salt out” 2,4-dinitrophenyl sulfate, but all
`these salts destabilize the transition state (Table I).
`The significant
`rate deceleration caused by sodium
`perchlorate (Table IV) is explicable in terms of a very
`pronounced transition-state destabilization which is
`not canceled by initial state interactions.
`It
`is not
`surprising that sodium chloride and sulfate affect the
`hydrolysis rate insignificantly since the former salt
`only slightly alters the molar activity coefficient of the
`initial and transition state and interactions of the latter
`with the initial state and transition state cancel.
`Copper (II) and silver(I)
`ions have no significant
`the hydrolysis of 2,4-dinitrophenyl sulfate
`effect on
`a considerable copper(II)
`(Table IV).23 Recently,
`reported for 8-hydroxyquinoline
`ion catalysis was
`sulfate.5 Although the exact nature of metal
`ion
`rate enhance-
`catalysis is not perfectly understood,
`ment is generally discussed in terms of chelation with
`the substrate7 or charge neutralization in the transition
`state.25 Not unexpectedly, metal ions do not seem to
`chelate with 2,4-dinitrophenyl sulfate nor do they en-
`hance its hydrolysis rate by any other mechanism.
`In their recent study Benkovic and Benkovic2 have
`reported that the rate constant for p-nitrophenyl sulfate
`hydrolysis was a factor of two slower in 50% aqueous
`acetonitrile than in water.
`these authors
`Indeed,
`used this fact as one of their mechanistic probes to
`substantiate the proposed unimolecular mechanism2
`since similar effects have been reported for hydrolysis
`of acetyl phosphate dianion.18 On the other hand, the
`hydrolysis rate of 2,4-dinitrophenyl phosphate was
`found to increase by a factor of four in aqueous 50%
`the
`acetonitrile,17 and considerable enhancement of
`hydrolysis rate of steroid sulfates28 by moist dioxane
`has been noted for
`time.
`These conflicting
`some
`effects of dipolar aprotic solvents on the hydrolysis of
`sulfate and phosphate esters question the validity of
`the use of solvent effects as mechanistic criteria, at
`least for these compounds. We,
`therefore, have ex-
`amined the effect of acetonitrile, dimethyl sulfoxide,
`dioxane, and  , -dimethylformamide on hydrolysis of
`2,4-dinitrophenyl sulfate (Table V). Changing the
`aqueous medium to aqueous acetonitrile steadily de-
`the hydrolysis rate. Dioxane has the same
`creases
`effect up to 80% dioxane concentration, but, when the
`water content of the medium is 10%, the neutral rate
`is increased by a factor of 70. Dimethyl sulfoxide and
` , -dimethylformamide increase the rate of the neu-
`tral hydrolysis exponentially.
`
`(23) However, the presence of 0.01 M cetyltrimethylammonium bromide,
`a cationic micelle, increases the rate of both the neutral and the base-cata-
`lyzed hydrolysis by a factor of two but very significantly decreases the acid-
`catalyzed rate of pH 2.0, whereas 0.01 M sodium lauryl sulfate, an anionic
`micelle, slightly decreases the rate of the neutral hydrolysis.54
`(24) E. J. Fendler and J. H. Fendler, to be published.
`(25) C. H. Oestreich and  . M. Jones, Biochem., 2926, 3151 (1966); 1515
`(1967).
`(26) J. McKenna and J. K. Norymberski, J. Chem. Soc., 3889 (1957).
`
`

`

`3858 Fendler and Fendler
`
`The effects of these dipolar aprotic solvents on the
`hydrolysis of 2,4-dinitrophenyl sulfate is revealed
`strikingly by their entropies of activation. Supported
`by a large body of experimentally determined entropies
`of activation for reactions of known mechanism, posi-
`tive AS* values have been assigned to unimolecular
`to bimolecular mecha-
`mechanisms and negative ones
`nisms.27 The entropies of activation for the neutral
`hydrolysis of nitrophenyl and dinitrophenyl sulfates
`are more negative than those generally associated with
`a unimolecular mechanism (Table III). The inter-
`pretation of the magnitude of entropy effects for uni-
`molecular reactions is, however,
`less straightforward
`than that
`reactions. The ordering
`for bimolecular
`effect in the latter case arises from the covalently bound
`in the transition state, while the former situa-
`water
`tion is a composite of two opposing effects. The one-
`step process of forming sulfur trioxide and phenoxide
`ion results in a decrease in the order of the system,
`hence an increase in the entropy. On the other hand,
`the internal mobility of the solvent
`is reduced by a
`greater solvation of the entities in the transition state
`resulting in a decrease in entropy. Apparently for
`aryl sulfate esters, the net effect on the entropy differ-
`ence between the initial and the transition state due to
`these opposing factors results in a moderatively nega-
`tive entropy of activation. Several reactions whose
`mechanisms have been independently established to be
`unimolecular have fairly considerable negative entro-
`pies of activation.28·29 The hydrolysis of 2,4-dinitro-
`phenyl sulfate in aqueous solvent mixtures containing
`dipolar aprotic solvents exhibits significantly higher
`entropies of activation (Table V). The effectiveness
`of the solvents in increasing the entropies of activation
`is dioxane >  , -dimethylformamide > dimethyl
`sulfoxide > acetonitrile. Tentatively one may ascribe
`the entropy increase to a decrease in transition-state
`solvation.28
`best described in
`Medium effects, however,
`are
`terms of eq 1 which allows a separation of the activity
`coefficient effects in the initial and transition states.
`We have attempted to do this for the acetonitrile and
`It
`is evident
`dimethyl sulfoxide systems (Table I).
`that, although acetonitrile considerably increases the
`molar activity coefficient of 2,4-dinitrophenyl sul-
`fate, the corresponding transiti

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket