Vous êtes sur la page 1sur 18

Kinetics of Alkaline Hydrolysis of Organic

Esters and Amides in Neutrally-Buffered


Solution
BRUCE A. ROBINSON
Los Alamos National Laboratory, Los Alamos, N e w Mexico

JEFFERSON W. TESTER
Massachusetts Institute of Technology, Cambridge, Massachusetts

Abstract
Reaction rate constants for the hydrolysis of organic esters and amides were determined a t
temperatures of 100-240C in aqueous solutions buffered at pH values between 5.5 and 7.3.
Experiments are modeled assuming alkaline hydrolysis with a thermodynamic solution model
included to account for the temperature dependence of hydroxide ion concentration. In most
cases, the ester hydrolysis second order rate constants agree well with published values from
experiments in strongly basic solutions a t pH values from 11 to 14 and temperatures from 25SOT, despite the large extrapolations required to compare the data sets. The amide hydrolysis
rate constants are about one order of magnitude higher than the extrapolated results from
other investigators, but the reaction rate increased proportionally with hydroxide ion concentration, suggesting that a n alkaline hydrolysis mechanism is also appropriate. These data establish t h e validity of t h e alkaline hydrolysis mechanism and can be used to predict
hydrolysis reaction rates in neutrally-buffered solutions such as many groundwater and geothermal fluids.

Introduction
The fate of organic compounds in neutrally-buffered fluids is important
in such diverse applications as transport of organics in groundwater, tracer
transport in geothermal reservoirs, and chemical enhanced oil recovery
techniques. Chemical reaction via hydrolysis or other reactions is one possible fate of an organic solute. Previous studies of hydrolysis reaction rates
in solutions buffered a t pH near neutral are scarce. Mabey and Mill [ l l
compiled alkaline hydrolysis data, usually obtained at high pH, for many
classes of organic compounds and extrapolated the results to groundwater
pH values. They claimed that extrapolations from higher pHs to obtain
upper bounds on chemical half-life at 298 K and pH 7 are justified.
The present study examines hydrolysis reaction rates of organic esters
and amides in solutions containing sodium bicarbonate and acetic acid.
This buffer system was chosen to provide a constant-pH solution in the
range of 5 to 8 while simulating a temperature dependence of pH similar
t o a natural groundwater fluid. The reactants were selected as possible
tracers for geothermal reservoirs 121. This provided the motivation to carry
out reaction kinetics studies at temperatures up to 240C. These reactants
may also be model compounds for possible degradation reactions of polyInternational Journal of Chemical Kinetics, Vol. 22, 431-448 (1990)
Q 1990 John Wiley & Sons, Inc.
CCC 0538-8066/90/050431-18$04.00

432

ROBINSON AND TESTER

mers used as viscosifying agents or in enhanced oil recovery techniques.


Finally, if agreement is obtained between alkaline hydrolysis data at neutral pH and elevated temperature and previous studies at much higher pH,
this will lend further justification t o the extrapolation of rate data t o
groundwater flow conditions. Of course, other effects such as possible catalysis due to the presence of mineral surfaces could be important and thus
must be examined.

Previous Studies
The class of ester hydrolysis reactions examined in this study is given
by:
(1)

RCOOR

+ H,O

RCOOH

+ ROH

The unsubstituted amide hydrolysis reactions treated in this study are represented by
(2)

RCONH,

+ H,O

RCOOH

+ NH,

The most comprehensive work on the ester and amide hydrolysis reactions
is that of Kirby [31. The rate of hydrolysis of most esters and amides is
strongly dependent on solution pH, which will be important when predicting
the reaction rate behavior in many groundwater and geothermal fluids.
The acid catalyzed, neutral hydrolysis, and alkaline hydrolysis mechanisms
may be combined in a n overall rate expression valid for any pH:
(3)

d [RCOOR]
= -h,,,[RCOOR]
dt

where
(4)

kObs= h,

+ K,[H+] + k,[OH-]

As we shall see, for geothermal fluids buffered in the pH range of 6-8,


the alkaline hydrolysis mechanism may be the important one:

-dC
_
dt

-k,[OH-]C

Although differences between the ester and amide hydrolysis mechanisms


have been identified [41, in the present study the data will be analyzed assuming eq. (5),unless it is obviously invalid. If [OH-] is kept constant during a n experiment, the hydrolysis rate is pseudo-first order in ester o r
amide concentration.
Varying the R and R groups changes the rate of reaction markedly in
some cases. Table I lists some Arrhenius parameters for various ester and
amide hydrolysis reactions [3-81, showing the effect of different R and R
groups. Full explanations of the influence of substituent groups on the rate
of alkaline hydroylsis of esters may be found elsewhere [6-101.

Chemistry of Neutrally-Buffered Solutions


Many naturally ocurring waters are buffered by the presence of dissolved
CO, and HC0,- ion, with the resulting pH ranging from approximately 5

433

KINETICS OF ALKALINE HYDROLYSIS

TABLEI. Published alkaline hydrolysis kinetic data.

log1oAr

Ea

Reaction Time
T ~ ( ~( )
s)

Ref.

Conditions

C6H5COOCH3

60% dioxane

34(b)

p-BrC H COOCH
6 4
3

60% dioxane

9(b)

CH3COOC2H5

70% acetone

6.10

42.3

20

CH3COO(CH2)2CH3

70% acetone

6.61

46.5

34

CH3COO(CH2)3CH3

70% acetone

6.70

47.3

39

CH3COOC( CH,, )

70% acetone

6.30

56.5

3960

CH3COOC2H5

water

7.39

47.3

C2H5COOC2Hs

water

6.21

44.9

45

11-C H COOCiiH5

water

6.15

42.7

21

( CH3)3CCOOC2H5

70% acetone

5.91

54.4

4160

CH3CONH2

water

5.99

59.5

2710

CH3(CH2)2CONH2

water

5.12

56.7

6.49~10~ 4

C6H5CONH2

water

5.76

61.0

8.42~10~ 5

3 7

(liter/mol-sec)

(kJ/mol)

Compound

"In 1 M NaOH a t 25"C, unless otherwise indicated.


Experiment performed a t 35C.

to 8. To perform hydrolysis kinetic experiments which simulate reaction in


a natural fluid, a buffer system with chemical content similar to the natural
fluid must be chosen. To simplify analysis of the experiments, the system
must have sufficient buffering capacity to keep the pH approximately constant even while organic acids are being produced from the hydrolysis of
ester or amide. Finally, the pH of the fluid must be adjustable over a range
of several pH units in the vicinity of pH 7.
The chemical buffer system chosen to accomplish these objectives was
the carbonate-acetate buffer, consisting of mixtures of dissolved sodium bicarbonate (NaHC0.J and acetic acid (CH,COOH). Mixtures of sodium bicarbonate and acetic acid in water will have near neutral pH, which makes
this system attractive for simulating geothermal brines of pH 5 to 8. Due to
the temperature dependence of the ionization equilibrium constants of water
and the buffer compounds, the hydroxide ion concentration, which appears
in the alkaline hydrolysis rate law, is much different than the value determined from the pH of a sample at room temperature. The two steps required
t o model this behavior for a kinetics experiment a t elevated temperature
are: (1)calculate the solution composition of all dissolved species a t 25C
after mixing sodium bicarbonate and acetic acid in water; and (2) determine

434

ROBINSON AND TESTER

the new equilibrium concentrations at the higher temperature. The calculations required for these steps are outlined below. The following equations,
given along with the equilibrium reactions they represent, must first be
satisfied at room temperature:

(6)

K,

[H][OH-]YHYOH; H,O

Hi

+ OH-

(7)

where Ac- is acetate ion, CH,COO-, and the ys are activity coefficients.
Two additional independent equations are the charge balance equation for
this system, and a mass balance on acetate:
(10)
(11)

[H] + a]

[OH-] - [HCO, ]
[HAc] + [Ac-]

[Ac-] - 2[CO,-]

[HAc],,

where IHAcI, is the total acetic acid concentration before dissociation. The
six equations listed above contain seven unknowns, one of which, [Hl, is
measured independently. These equations are largely uncoupled, and thus
can be solved easily.
When the solution is heated to temperature T, the compositions will change
from their values a t room temperature because of the temperature dependence of the equilibrium constants and activity coefficients in eqs. (61-49).
For example, Figure 1 is a plot of Kw as a function of temperature [111.
Correlations for other ionization equilibrium constants were obtained for
K,, the ionization of acetic acid, and K, and K3, the ionization constants of
carbonic acid [12-131. Once the vessel is closed and the temperature is
raised, only the equilibrium reactions given in eqs. (6)-(9) above are possible, and these proceed until the new equilibrium constants are obtained
at the elevated temperature. To calculate the concentrations a t temperature T, four unknowns are defined representing the mols per liter of H ion
produced by dissociation in the four reactions. The resulting equilibrium
constant expressions then contain only these four unknowns, which are
solved simultaneously to obtain the concentrations.
Single-ion activity coefficients were calculated using a n extended DebyeHuckel model [141:
.log yt

Az p I
1 -t aJ312

where yi is the activity coefficient, I is the ionic strength, z, is the charge


on the ion, and A, B , and a, are the parameters of the Debye-Huckel model.
Of course, single-ion activity coefficients cannot be obtained directly: only
their products can be measured. However, in hydrothermal fluids with I up
to 1 m, eq. (12) applies, with solvent parameters A and B tabulated as a

435

KINETICS OF ALKALINE HYDROLYSIS

50

I
K,

Temperature ("C )
I00 125 150 200 2 5 0 300

75
I

[H+][OH-]

,.+.\

10

a,,-@

K w 10

i
f

10

10

/.

FIscher and B a r n e s ( 1 9 7 2 )

Figure 1. Ionization equilibrium constant for water as a function of 1/T

function of temperature and values for a, given for individual ions [14]. A t
the relatively low ionic strengths of about 0.05 m of interest in the present
study, we also assume that the activity of water and the activity coefficients of undissociated carbonic acid and acetic acid are equal to unity.
Figure 2 shows the calculated concentration of OH- as a function of temperature for typical buffer concentrations. The model predicts [OH-] to be
a strong function of temperature. Thus, hydroxide ion concentration must
be calculated for each experiment when using the alkaline hydrolysis
mechanism to model the ester and amide hydrolysis kinetic data. The simplistic assumption that [OH-] = lO(pH,,-'*), where pH,, is the value measured at room temperature, is clearly incorrect for this buffer system. A
similar behavior is also found for neutrally-buffered hydrothermal fluids
[ E l , making the carbonate-acetate buffer a n appropriate model solution
for performing kinetics studies.

Experimental
The ester and amide hydrolysis kinetic experiments a t elevated temperature were performed in a one gallon (3.78 liter), Pressure Products
stainless steel stirred autoclave reactor with automatic temperature con-

436

ROBINSON AND TESTER

10-3

10."

y
-

10-5

10-6

I
L?

10-7

10-0

10-9

20

60

100

140

180

220

260

300

340

TEMPERATURE ( O C I

Figure 2. Calculated value of hydroxide ion concentration versus temperature for


typical buffer conditions: LNaHC0310 = 4.23 x
mol/liter, [HAc], = 4.52 x
moI/Iiter.

trol (k2"C). About 3.5 liters of buffered solutions containing the ester or
amide were charged to the reactor a t room temperature and pressure. The
sealed vessel was then heated to the desired temperature. Experiments a t
temperatures above the boiling point of water were operated at the vapor
pressure of water at that temperature. This allowed samples to be expelled
through hypodermic tubing by simply opening a valve connected t o the
sampling port. Due to the density change of water with temperature, the
fluid typically expanded to fill the entire vessel, as evidenced by a rapid rise
in pressure above a certain temperature during heatup. Some fluid was
bled off to accomodate this expansion. Thus, the vapor space in the reactor
at temperature was small or nonexistent, and partitioning of the reactants,
products, and buffer compounds into the vapor may be neglected.
For the kinetics experiments carried out below 100C, a n inert atmosphere of nitogen (about 50 psi) was kept over the solution to enable sampling in a similar manner. A length of coiled tubing was immersed in a
cold water bath to cool the sample to room temperature during sampling.
Usually, 8-10 samples were collected during the course of a n experiment
to determine its progress as a function of time. Samples which could not be
analyzed immediately were refrigerated to prevent further reaction.
Initial concentrations ranging from about 50-200 ppm of reactant were
typically used, and the conversion was typically 50-90%, with a few experiments terminated a t lower concentrations. To adjust the pH of the solution,
the concentration of sodium bicarbonate was held constant a t 0.0328m,
and the concentration of acetic acid was varied. Experiments were carried

KINETICS OF ALKALINE HYDROLYSIS

431

out at three different acetic acid concentrations: a 0.350m solution resulted


in a pH of about 5.6, a 0.219m solution gave a pH of about 6.5, and a
0.00656m solution gave a pH of about 7.3.
The ester hydrolysis reactants and products were measured using a
Hewlett-Packard Model 58308 gas chromatograph with a flame ionization
detector (FID). Liquid samples were injected with a hypodermic syringe a t
a temperature high enough to vaporize the water, ester, and alcohol. A
1 m long Porapak Q column was used, which produced sharp, nonoverlapping peaks for the alcohol and ester. We estimate the error of these measurements to be about +5%. The responses of the organic acids were erratic in
both the samples and standard solutions, so these results were not utilized
in the reaction calculations. The progress of reactant concentration was
followed over time with the concentration of the alcohol product measured
to check for consistency in the mass balance. Typically, the mass balance
agreed a t least to within +-lo%.For the unsubstituted amides tested, reaction progress was monitored by measuring the reaction product ammonia
(NH,) using a n Orion ammonium ion-specific electrode, with an accuracy
of about 55%.Finally, the solution pH a t room temperature was measured
with an Orion pH electrode, with an accuracy of 20.02 pH units.
In addition to the experiments in neutrally-buffered solutions a t elevated temperatures, several experiments were carried out a t higher pH
and lower temperature for comparison. Aqueous solutions of the ester were
prepared and heated to the desired temperature in a stirred glass beaker.
At t = 0, an aliquot of concentrated NaOH solution was injected. With the
ester as the excess reactant, reaction progress was monitored by measuring [OH-] using a pH electrode. Typically, a linear ln[OH-l versus time behavior was observed over at least one pH unit, thus verifying the alkaline
hydrolysis rate law under these conditions. Detailed description of these
experiments may be found elsewhere 1151.

Results and Discussion


Table I1 lists the first order rate constants for each isothermal experiment in carbonate-acetate mixtures a t elevated temperature. Also shown
are the calculated hydroxide ion concentration and calculated second order
rate constant. Table I11 summarizes the results of beaker experiments at
low temperatures and elevated pH. In the discussion below, we address the
validity of the alkaline hydrolysis mechanism and OH- model with our kinetics results in neutrally-buffered solutions by comparing these results to
published alkaline hydrolysis kinetics data and the beaker experiments of
the present study. In most cases, we show the alkaline hydrolysis rate law
t o be valid; second order Arrhenius parameters for the reactions and error
estimates for the activation energies are listed in Table IV. Uncertainties
in the rate constants and activation energies were computed using a Monte
Carlo anlaysis in which random errors were placed on the concentration
and pH measurements. Many realizations of a n experiment were performed and the standard deviation of the calculated values were computed.
Error estimates were obtained assuming a 5% error on concentration and

438

ROBINSON AND TESTER

TABLE11. Summary of ester and amide hydrolysis kinetic experiments.

T('C)

ISt order
rate const.
k (s-')

PH
(at 25'C)

[OH-]
calculated
at T ( M )

Znd Order
rate const.
k2
(liter/mol-s)

Ethyl Acetate (EA) CH3COOC2H5


110
131
132
150
172
190

3.45~10-~ 5.23~10-~
4 . 9 0 ~ 1 0 -+_~ 1.56~10-~
1 . 4 9 ~ 1 0 -f~ 1.22~10-~
5 . 5 4 ~ 1 0 -_+~ 3 . 9 0 ~ 1 0 - ~
2 . 2 7 ~ 1 0t~ 1.42~10-~
~
3 . 3 2 ~ 1 0 -+~ 1 . 3 5 ~ 1 0 - ~

5.64
7.28
5.52
5.60
5.55
5.47

6.54xw7

5.27 t 0.974
4 . 6 7 ~ ~ ~ 10.5 f 0.860
1. 22x10-6
12.3 f 1.49
2.96x10@
18.7 f 2.18
5.78~10~~
39.2 f 4.50
8.85xd
37.5 f 4.04

Ethyl Propionate (EP) C2H5COOC2H5


108
135
153
173
191

1 . 7 5 ~ 1 0 -+~ 2 . 7 6 ~ 1 0 - ~
9 . 7 0 ~ 1 0 -+~ 6.36x1C7
3 . 2 5 ~ 1 0 -+~ 3.94~10-~
1.17~10-~ 1.27~10-~
3 . 5 8 ~ 1 0 -+~ 1.5i~io-~

5.53
5.46
5.48
5.52
5.47

4.53x10-?
1.18~10-~
2.48~10-~
5.58~10-~
9.15~10-~

3.86

8.20

+ 0.883

0.691

1.99
3.10
39.1 t 4.26
13.1

21.0

Isopentyl Acetate (IPA) CH3COOCH2CH2CH(CH3)CH3

118
143
182

3.45~10-~ 5.65~10-~ 5.59


2 . 3 4 ~ 1 0 -2~ 1 . 4 0 ~ 1 0 - ~ 5.57
5.56
2.99~10-~ 1.20~10-

8. 14x1C7

2. 12x10-6

8.32~10-~

4.24 f 0.807
11.1 f 1.22
35.9 f 3.74

Hexyl Acetate (HA) CH3COOC6H13


116
131
140
153
153
171
181

f 1.42~10-~
5.08~10-~
8.87Xi0-~ f 7.88~10-~
1.50~10-~ 1.56~10-~
4 . 9 5 ~ 1 0 -f~ 4 . 4 0 ~ 1 0 - ~
2 . 1 O ~ l O -f~ 1 . 8 5 ~ 1 0 - ~
1 . 0 6 ~ 1 0 - ~ 1.63~10-~
2 . 2 7 ~ 1 0 - ~ 7.57~10-~

7.16
5.48
5.53
5.54
6.42
5.53
5.62

2.05~10-~
1.06~10-~
1.71~10-~
1.90~10-~
1.90~10-~
5.33~10-~
9.24~10-~

2.48 + 0.732
8.37 f 1.03
8 . 8 0 f 1.20
17.3 f 2.31
11.0 + 1.40
19.9 f 3.64
24.5 f 2.45

Ethyl Pivalate (EPI) (CH3)3CCOOC2H5

176
178
194
217
232

2 . 4 9 ~ 1 0 -2~ 2.08~10I. ~ 1 ~ 1 0 -1.89~10~


1.36~10-~
2 2.06~105.17x1C5 _+ 6 . 2 8 ~ 1 0 - ~
2 . 3 8 ~ 1 0 -f~ 4.30~10-

5.56
7.23
5.56
5.52
5.54

6.79~10-~
2.08~10-~
1.24~10-~
2.35~10-~
3.8 8 x ~ ~ - 5

0.367 + 0.0466
0.582 f 0.0964
1.10 +_ 0.199
2.20 t 0.355
6.14 t 1.34

439

KINETICS OF ALKALINE HYDROLYSIS

TABLE11. (Continued)
t-Butyl Acetate (TBA) CH3COOC(CH3)3

62.5
78
102
106

2.95~10-~
f 1.85~10-~
2.63~10-~ 1.67~10-~
2. 36xw4
2. 6 5 x W 5
3.36~10-~
f 4.01~10-~

5.50
5.50
5.48
6.35

Acetamide (AA) CH3CONHZ

157
161
170
173
204
223

1.04~10-~
f 9.91x10-'
+ 5.66~104.26~10-~
2.46~10-~
f 2.13~10-~
1.79x1C5 2 1.74~10+ 8.11~10-~
5.26~10-~
1.32~10-~
f 6.00~10-~

5.52
7.23
5.60
6.38
5.61
5.57

3.16~10-~
1.23~10-~
6.07~10-~
3.75xuY5
1.~ O X ~ O - ~
3.14~10-~

0.329 k 0.0437
0.345 f 0.0486
0.405 f 0.0523
0.477 f 0.0626
2.77 f 0.510
4 . 2 0 f 1.98

Butyramide (BU) C4H9CONH2

170
173
180
186
187
202
206
213
222

223
2 38

f 3.85~10-~
3.82~10-~
2.51~10-~ 2.21x10-7
7
f 3.34~103.57~10-~
f 5.O4xW7
5.77~10-~
6
3.54x1C5 3.09~101.WXIO-~ 2 1.91~10-~
f 2.32~10-~
2.44~10-~
6
4.25~10-~
+ 4.71~104.68~10-~
f 5.36~10-~
6
6.~ ~ x I O 7.21~10- ~
5
1.45~10-~
2 1.47~10-

7.31
5.65
5.60
5.60
6.65
5.66
5.61
5.66
5.69
5.60
5.71

1. ~ O X ~ O - ~
7.57~10-~
8.53~10-~
1.04~10-~
1.O ~ X ~ O - ~
1.99~10-~
2.03~10-~
2.79~16~
3.
3.34~10-~
6.40~10-~

+ 0.0242
0.332 f 0.0434
0.418 + 0.0552
0.554 f 0.0726
0.325 + 0.0388
0.965 +_ 0.135
1.20 + 0.167
1.52 + 0.228
1.21 2 0.177
2.05
0.305
2.27 + 0.272
0.212

Benzamide (BZA) C6H5CONH2


171

184
210

222

+ 2,30x102.92~10-~
+ ~5.92x1C7
6.41~10
~
2 1.93~10-~
2.Ol~lO-~
k ~6.54~10-~
6.82~10
~

5.66
5.64
5.61
5.65

7.24~10-~
1.07~10-~
2.29~10-~
3.~ ~ x I O - ~

0.403 f 0.0507
0.600 f 0.0935
0.877 f 0.123
1.90 f 0.267

1~0.02pH units, and the values in the tables are for two standard deviations, representing a 95% confidence interval.
Ethyl Acetate: The kinetics of ethyl acetate hydrolysis have been studied
extensively, making this hydrolysis reaction the best choice for testing the
validity of the alkaline hydrolysis mechanism and [OH-] model developed
previously. Two independent checks of the model are possible. First, second
order alkaline hydrolysis kinetics parameters from the experiments carried out at similar buffer conditions can be compared to published kinetics
results and the beaker experiments listed in Table 111. Figure 3 shows that

440

ROBINSON AND TESTER

TABLE111. Second order alkaline hydrolysis rate constants for the beaker
experiments.
T(OC)

k2(liter/mol-s)

27

0.123

32

0.160

36

0.232

40.5

0.295

46

0.432

50

0.476

55

0.667

60

0.840

Isopentyl Ace a

65

0.307

Ethyl Propionate

45

0.246

Ester

Ethyl Acetate

the calculated second order rate constants k, agree well with those of the
beaker experiments and other studies [ 16-17]. The agreement is impressive considering the large range of extrapolation of both temperature and

TABLEIV. Arrhenius parameters for the ester and amide hydrolysis reactions.
Compound

(liter/molLsec)

Ea (kJ/mol)

Ethyl Acetale

1.01x106

38.4t3.0

Ethyl Propionate

6
1.18x10

40.2f3.9

Isopentyl Acetate

1.65~10'

49.3t5.1

Hexyl Acetate

6
8.36~10

47.5k4.7

Ethyl Pivalate

2.44~10~

83.8t7.7

t-Butyl Acetate(a)

1.36~10'~

Acetamide

8
4.54~10

76.0t10.8

Butyramide

2.25~10'

67.6k3.7

Benzamide

5
2.47~10

49.3k6.9

"First order rate constant, A , in units of K 1

113t2.7

441

KINETICS OF ALKALINE HYDROLYSIS

'I1l.i I

0lJi:i

!)Il'i

0027

111) 2 s

!I0 ( I

I)!) I $

I,,: i l

I . 'I' (I\ 'i


Figure 3. Second order rate constant for ethyl acetate: comparison of buffer and beaker
experiments. All low-temperature experiments were carried out in pure water with
various NaOH concentrations.

pH. For example, the beaker experiments were pseudo-first order reactions
with ethyl acetate in excess, in the temperature range of 25 to 60C and
pH of 11, compared to those carried out at a room-temperature pH of 5.6
and temperatures from 110-190C in the buffer experiments. By contrast,
without considering the temperature
had we assumed [OH-] = 10(pH25-14),
effect on K , and hydroxide concentration, the calculated values of k, from
the buffer experiments would be 2 to 3 orders of magnitude too high. The
[OH-] model properly corrects this discrepancy and accounts for the larger
apparent activation energy of the first order rate constants. Almost half of
the observed activation energy can be attributed t o the increase in [OH-]
with temperature due to the water dissociation reaction.
An independent test of the model is to measure the rate behavior for a
change in solution pH by varying the acetic acid concentration. The variation in room-temperature pH of the buffer solution from 5.6 to 7.3 resulted
in an appropriate increase in hydrolysis rate. The reaction rate is first order in [OH-] to within experimental accuracy, resulting in similar values
of the calculated second order rate constants.
Ethyl Propionate: In an alkaline hydrolysis beaker experiment a t 45C
starting a t pH 11 in excess ethyl propionate, k, was determined to be
0.673 literimol-s. Figure 4 shows close agreement with the autoclave data
despite the large extrapolation. Alkaline hydrolysis data in 70% acetone
solutions [61 are also shown in the figure. The agreement is adequate, suggesting that the buffer experiment data are modeled quite well with the alkaline hydrolysis mechanism and [OH-] model.

442

ROBINSON AND TESTER

110:21

001' I

0025

(10,:!I

(1027

111'

(K

00 I 1

illl.! !

;\

Figure 4. Second order rate constants for ethyl propionate. The experiments of Davies
and Evans 161 were carried out in pure water.

Isopentyl Acetate: Figure 5 compares extrapolated buffer experiment data


with a beaker experiment at 55C. The agreement is excellent, suggesting
that the alkaline mechanism and [OH-] model are again valid for hydrolysis
in the buffer solution.
Hexyl Acetate: Literature data were unavailable for the alkaline hydrolysis of hexyl acetate. Furthermore, the beaker experimental technique of
measuring lz2 in an excess of the ester was not possible because of the low
solubility of hexyl acetate in water. Instead, in Figure 6, the comparison is
made with kinetics data for n-butyl acetate in 70% acetone solutions [7].
They studied steric effects of branched and straight-chained substituents
on the alcohol side of the ester, concluding that for esters of normal (unbranched) alcohols, the addition of each carbon to the chain has a lesser effect on reactivity. This means that the hydrolysis rate of n-butyl acetate
should be comparable to that of hexyl acetate, which simply has two more
unbranched carbons several atoms away from the carbonyl reaction site.
The extrapolated n-butyl acetate line agrees quite well with the hexyl acetate data.
Two buffer experiments a t different values of room temperature pH (6.42
and 7.16) were also performed for hexyl acetate. The reaction rates of these
two experiments were 4.2 and 30 times faster than would have occurred at
room temperature pH of 5.6. These increases are proportional to the increase in [OH], and thus the second order rate constants agree well with
values a t lower pH. This evidence, along with agreement with the cited n-

443

KINETICS OF ALKALINE HYDROLYSIS

3 Buffer
?

Experimenls

Beakpr E x p e r i m e n t s

01

o m

0021

0015

0027

0029

J3 I

00 I1

'

T (I\ 1

Figure 5 Second order rate constants for isopentyl acetate. The low-temperature experiment was carried out in pure water with NaOH added

liC21

I"'",

00235

00245

O(1L 53

I/,, & - : j
Figure 6. Second order rate constants for hexyl acetate. The n-butyl acetate experiments of Jones and Thomas [71 were carried out in 70% acetone solutions.

444

ROBINSON AND TESTER

butyl acetate results [71, suggests that the alkaline hydrolysis mechanism
and [OH-] model are also appropriate for hexyl acetate.
Ethyl Piualate: Figure 7 compares the buffer experiments to the extrapolated data for ethyl pivalate alkaline hydrolysis cited by Kirby [3]. The
agreement in the temperature range of the autoclave experiments is close,
but this may be fortuitous considering the difference in activation energy
of the two sets of data. Experimental uncertainty could explain this discrepency. Erratic gas chromatograph results and solubility and volatility
problems may have resulted in larger data scatter. Low solubility precluded the use of our technique for measuring alkaline hydrolysis rate in
excess ester. One data point obtained at a room-temperature pH of 7.23 appears to be governed by the alkaline hydrolysis rate law. This experiment
supports the validity of the [OH-] and alkaline hydrolysis mechanism for
ethyl pivalate.
t-Butyl Acetate: This is the only organic ester of the six studied which
does not obey the alkaline hydrolysis mechanism in any respect. Previously determined values of k, for alkaline hydrolysis [3] fall several orders
of magnitude below those calculated from the buffer experiments. Moreover, in Experiments TBAl and 2, a change of starting pH which should
have increased the observed rate by a factor of 7.5 had no effect. The rate
appears to be independent of pH, suggesting that in this range of pH the
neutral hydrolysis mechanism is operative. Since the hydrolysis of t-butyl
acetate and other esters of tertiary alcohols becomes acid catalyzed at

Figure 7. Second order rate constants for ethyl pivalate. The experiments reported by
Kirby [31 were carried out in 70% acetone solutions.

445

KINETICS OF ALKALINE HYDROLYSIS

much higher values of pH [3], the transition pH for the alkaline hydrolysis
mechanism is probably higher as well.
Amide Hydrolysis Experiments: As illustrated in Figures 8-10, buffer
experiments for the three organic amides tested iacetamide, butyramide,
and benzamide) result in alkaline hydrolysis k, values about an order of
magnitude higher than the extrapolated lines obtained for acetamide 141,
butyramide [4], and benzamide [51. However, the dependence of the rate on
room-temperature pH is modeled accurately assuming alkaline hydrolysis
for acetamide and butyramide (this effect was not examined for benzamide). Thus, in this sense the alkaline hydrolysis rate law is consistent
with the data
Of course, lack of agreement between the buffer experiments and the
results of previous investigators casts some doubt on the validity of the proposed mechanism. The most likely explanation is that a different type of
alkaline hydrolysis mechanism with the same overall rate expression but
different kinetics parameters may be occurring under these conditions. Alternatively, the [OH-] model ias opposed to the alkaline hydroxide mechanism itself) could be inadequate a t the high temperatures a t which these
experiments were carried out. Possibly, [OH-] deviates from the calculated
values a t high temperature while still exhibiting the correct trend with
room temperature pH. Finally, the extent of extrapolation might be too
large given the uncertainty in experimentally-determined rate constants.
Possible Catalysis B y Mineral Surfaces: To test for the possibility of catalysis of the reactions, sealed titanium bombs containing a sample of

01
002

0021

0022

1/T

002 3

00'4

IK '1

Figure 8 Second order rate constants for acetamide The experiments of Bruylants
and Kezdy 141 were carried out in pure water.

446

ROBINSON AND TESTER

0019

002

0023

0021

Figure 9. Second order rate constants for butyramide. The experiments of Bruylants
and Kezdy [4] were earried out in pure water.

0019

00'

00'1

OW?

032 i

I 'l'(h '\

Figure 10. Second order rate constants for benzamide. The experiments of Bender
were carried out in pure water.

[5]

KINETICS OF ALKALINE HYDROLYSIS

447

crushed granite and the reactant in a buffered solution were agitated in a


constant temperature oven. The granite used was a sample of the Carnmenellis granite obtained from a geothermal reservoir in Cornwall, U.K.,
since the primary motivation of this study was to identify appropriate tracers
for geothermal reservoirs. This granite contains mostly alkali feldspars and
quartz, along with significant quantities of fine clay particles. All compounds except hexyl acetate and t-butyl acetate were examined by running
experiments at identical conditions except for the presence or absence of
crushed granite. Close agreement was obtained between the second order
rate constants with or without granite present, suggesting that the reactions are not affected by the presence of mineral surfaces.

Conclusions
The kinetics of hydrolysis of most organic esters and amides tested were
found to follow an alkaline hydrolysis rate law at elevated temperatures in
solutions buffered at pH values between 5.5 and 7.3. The solutions were
buffered with sodium bicarbonate-acetic acid mixtures to prevent the product organic acids from affecting the pH. The concentration of [OH-] is a
strong function of temperature due to the temperature dependence of the
equilibrium constant for water dissociation. Similar dependence is found in
natural groundwater fluids containing bicarbonate ion and dissolved COz,
making this buffer an appropriate analog for studying hydrolysis in natural fluids. When the temperature effect is taken into account, calculated
second order rate constants for most ester hydrolysis reactions measured in
the present study agree well with values from experiments in strongly basic solutions at lower temperature. Also, adjusting the hydroxide ion concentration by changing the acetic acid concentration in the buffer yielded a
proportional change in reaction rate, as predicted by the alkaline hydrolysis mechanism. Ethyl acetate, ethyl propionate, ethyl pivalate, hexyl acetate, and isopentyl acetate all exhibited this behavior. Only t-butyl acetate
did not, possibly due to a shift to a neutral hydrolysis mechanism under
these conditions.
The amide hydrolysis reactions showed a change in reaction rate linearly proportional to [OH-], implying an alkaline hydrolysis mechanism.
However, the second order rate constants were about an order of magnitude larger than values extrapolated from results of other investigators in
experiments in more basic solutions at lower temperatures. Possible causes
of the discrepency are: (1)a different type of alkaline hydrolysis mechanism with different overall rate; (2) inadequacies in the buffer solution
model, resulting in an underestimation of [OH-] at high temperatures; or
(3) error caused by too large an extrapolation of experimental data from
one set of conditions t o another. Nonetheless, we have established the
validity of the alkaline hydrolysis rate law at elevated temperatures in
neutrally-buffered solutions for all of the compounds tested except t-butyl
acetate. By accounting for the strong temperature dependence of [OH-] in
solutions of pH near neutral, hydrolysis rates in natural groundwater or
other aqueous solutions can be predicted using data cited here. Furthermore, the presence of mineral surfaces does not appear to affect the reac-

448

ROBINSON AND TESTER

tion rates. Thus the present study further justifies the claim [ l l t h a t
alkaline hydrolysis kinetics can be accurately extrapolated to groundwater
chemical conditions.

Acknowledgment
Partial support was provided by the U.S. Department of Energy, Geothermal Technology Division. We also acknowledge the technical support offered
by Lois Gritzo, Lee Brown, Chuck Grigsby, Steve Birdsell, and Ron Aguilar.
Hari Viswanathan performed the Monte Carlo error analysis calculations.

Bibliography
[ll W. Mabey, and T. Mill, J . Phys. Chem. Ref. Data, 7,2 (1978).
[21 B. A. Robinson, J. W. Tester, and L. F. Brown, SPE Reservoir Eng., 28,3 (1988).
[31 Kirby, in Comprehensive Chemical Kinetics 10. Ester Formation and Hydrolysis, Related
Reactions, C. H. Bamford and C. F. H. Tipper, Eds. Elsevier, Amsterdam, 1972.
I41 A. Bruylants and F. Kezdy, Rec. Chem. Prog., 21 (1960).
[51 M. L. Bender, R. D. Ginger, and J. P. Unik, J . Am. Chem. Soc., 80 (1958).
[61 G . Davies and D. P. Evans, J . Chem. SOC.,(1940).
[71 R. W. A. Jones and J. D. R. Thomas, J . Chem. SOC.(B), (1966).
[81 C. K. Hancock and C. P. Falls, J . Am. Chem. Soc., 83 (1961).
[91 R. W. Taft, J . Am. Chem. Soc., 74 (1952).
[lo] D.F. DeTar and C. J. Tenpas, J . Am. Chem. Soc., 98 (1976).
[ l l ] J.R. Fisher and H.L. Barnes, J . Phys. Chem., 76 (1972).
[12] C. S. Patterson, G. H. Slocum, R. H. Busey, and R. E. Mesmer, Geochim. et Cosmochim.
Actu, 46 (1982).
[13] F. R. Bichowsky and F. D. Rossini, The Thermochemistry of the Chemical Substances,
Reinhold, New York, 1936.
[14] R. W. Henley, A. H. Truesdell, and P. B. Barton, Fluid Mineral Equilibria i n Hydrothermal Systems, Vol. 1, Society of Economic Geologists, El Paso, 1984.
[15] B. A. Robinson, Non-Reactive and Chemically Reactiue Tracers: Theory and Applications;
PhD Thesis, MIT, Cambridge, Massachusetts, 1985.
[16] R. David and J. Villermaux, Znd. Engng. Chem. F., 19 (1980).
[171 A. A. Frost and R. G. Pearson, Kinetics and Mechanism; John Wiley and Sons, New York,
1965.

Received December 12, 1988


Accepted October 23, 1989

Vous aimerez peut-être aussi