Vous êtes sur la page 1sur 7

View Article Online / Journal Homepage / Table of Contents for this issue

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

Inorg. Phys. Theor.

1747

Kinetics ofithe Bromate-Iodide Reaction : Catalysis by Acetate and Other


Carboxylate Ions
By A. F. M. Barton t and

G. A. Wright,' Department of Chemistry, University of Auckland, Auckland, New

Zealand
A direct amperometric method for continuous determination of iodine concentrations of the order of 1O-6M was
used to investigate the kinetics of the bromate-iodide reaction. Marked catalysis by buffer anions was discovered.

At 25' and
The rate law was found to be v = -d[BrO,-]/dt= ko[H+]2[Br0,-J[l-] + kb[B-][H+]a[BrO,-][l-].
ionic strength 1 .OOM in a perchlorate medium, k, = 49 M-, sec.-l, and for B- = acetate, P-chloropropionate, and
chloroacetate, kb = 5.4 x lo4, 1.7 x 1 04, and 2.5 x lo3 M-* sec.-l respectively. For acetate solutions more
concentrated t h a n 0.05M the second term predominated. The mechanism is discussed, and it is shown that the
intermediate Br02+ is unlikely to exist. Reaction schemes are proposed involving additions and replacements on
the bromate ion.

THEreaction between bromate and iodide is well known


in the iodometric method for bromate determination,
BrO,-

+ 91- + 6H+ --t Br- + 31,- + 3H,O

Stoicheiometric investigations were made by Randall


and Britton and Britton,, and the reaction has been
studied ~alorimetrically.~Tolstikov and Kolthoff and
Belcher 5 have discussed the kinetics with regard to the
analytical use of the reaction. It has been found that a
catalyst such as molybdate can be used to speed up the
reaction for a n a l y s i ~ . ~
Ostwald 6 in 1888 found that the reaction was accelerated by acids. Further studies were aimed at formulating the retarding action of the product iodine 7 and the
effect of acids, salts, and organic compounds.6.8 Noyes 9
proved that the rate was proportional to [H+I2. The
first determination of a full rate law was made by
Clark,lo measuring the initial rate.
ZI

-d[BrO,-]fdt

= Qd[I,-]/dt
= ko[H+]2[Br0,-] [I-]

Bowman 11 investigated the oxidation of arsenious acid


in bromate-iodide solutions due to the reaction with
iodine,
H,AsO,
I, H,O ----t H,AsO,
2HI
The rate equation (1) was confirmed in the presence of
arsenious acid to remove iodine, by Skrabal and
Schreiner.l2 Abel and Fabian l3 investigated the hydrogen isotope effect. Sigalla l4 considered, and rejected,
the possibility of a free radical mechanism in the bromateiodide reaction. He showed that, if a reaction involves
iodine atoms as intermediates, then autoxidation will
be induced. But the amount of iodine liberated in the
bromate-iodide reaction conforms to the same stoicheiometric equation in the presence and absence of dissolved
oxygen. One-electron transfers and radicals are found
only in the presence of ionizing radiation when the
hydrated electron and hydrogen atom are reducing
agents.15 Indelli et a1.l6J7 studied the bromate-iodide

+ +

(1)

-f Present address : Chemistry Department, Victoria University

of Wellington, New Zealand.


1 D. L. Randall, J . Amer. Chem. SOC.,1910, 32, 644.
2 H. T. S. Britton and H. G . Britton, J . Chem. SOC.,1952,
3887.
3 H. C. Mel, U.S. Atomic Energy Comm. URCL 2330, 1953,
2 (Chem. Abs., 1954, 48, 6228).
4 V. P. Tolstikov, Sbornik Statei Po obshchei Khim., Akad.
Nauk S.S.S.R., 1953, 2, 1249 (Chem. Ab;., 1955, 49, 2921).
5 I. M. Kolthoff and R. Belcher,
Volumetric Analysis,'
Interscience, New York, 1957, vol. 111, p. 269.
6 W. Ostwald, 2.phys. Chem., 1@38, 2, 127.
W. Meyerhoffer, 2. phys. Chem., 1888, 2, 585.

0. Burchard, 2. phys. Chem., 1888, 2, 796; G. Magnanini,

Gazzetta, 1890,20, 377; N. Schiloff,2. phys. Chem., 1898, 27, 513;

R. H. Clark, J . Phys. Chem., 1907, 11, 353.


-a A. A. Noyes and W. 0. Scott, 2.phys. Chem., 1895,18, 118;
A. A. Noyes, ibid., 1896, 19, 599.
lo R. H. Clark, J . Phys. Chem., 1906, 10, 679.
11 F. C. Bowman, J . Phys. Chem., 1907, 11, 292.
l2 A. Skrabal and H. Schreiner, Monafsh., 1935, 65, 213.
l3 E. Abel and F. Fabian, Osterr. Chem.-Ztg., 1937, 40, 26;
Monatsh., 1938, 71, 153.
14 J. Sigalla, J . Chim. phys., 1958, 55, 758.
l5 M. Anbar and P. Neta, J . Inorg. Nuclear Chem., 1966, 28,
1645.
16 A. Indelli, G. Nolan, and E. S. Amis, J . Amer. Chem. SOC.,
1960, 82, 3233.
1 7 A. Indelli, J . Phys. Chem., 1961,65, 240,972; A. Indelli and
G. Mantovani, Trans. Faraday SOC.,1965, 61, 909.

View Article Online

J. Chem. SOC.(A), 1968

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

1748
reaction because of its suitability for an investigation
of kinetic salt effects. The transition state corresponding to the rate law (1) has zero charge and the reactants
each have unit charge, so deviations from the DebyeHiickel law were not expected to start until concentration levels higher than those found in the case of multivalent ions. The extended form of the Bronsted-Debye
equation was obeyed in the presence of inert salts.
Several mechanism schemes have been proposed for
this and the other halate-halide reactions, but in general
evidence is inadequate to support one particular reaction
mechanism.18 The present investigation, using a direct
amperometric method for iodine determination, has
revealed marked catalysis by carboxylate buffers which
provides more mechanistic information.

by means of a recording potentiometer. A 50 PF capacitor


was included in parallel with the potentiometer to minimise
electrical noise. The determination was carried out in the
presence of bromate, perchlorate, and oxygen ; although
the E, values for the reduction of these compounds are
greater than the tri-iodide reduction potential, the reactions
are slow 23 and do not interfere. The proportionality constant between the measured current and the iodine concentration depended on the particular iodide concentration,
cell geometry, and stirring speed, and i t was determined
individually for each kinetic run by adding portions of
)
a micro-burette. A
standard iodine solution ( 0 . 0 0 3 ~with
diagram of the chart record for a typical experiment is
shown in Figure 1. The reaction rate ZI defined by equation
( 2 ) was measured from the recorder chart, and the extent

EXPERIMENTAL

Direct Ampevometric Iodine Analysis.-An


initial rate
method of kinetic investigation was used, the concentrations
of bromate, iodide, hydrogen ion, and carboxylate being
varied in successive runs to determine the order with respect
to each reactant in the rate law. The combined requirements of low concentration reactants to make the rate slow
enough to be measurable, and a small change in concentrations for the initial rate method meant that small concentrations of iodine, 10-6-10-5~, had to be determined
quantitatively during the kinetic runs. Many electroanalytical methods are available for iodine l9 but Potter 20
has shown that amperometry is the most sensitive. The
amperometric circuit in the present method 2 1 acts as an
electrochemical transducer giving an electrical signal which
varies in a linear manner with the iodine concentration.
When a steady state is reached for a fast electrochemical
reaction a t an electrode in a stirred solution containing a
base electrolyte, the electrolysis current becomes diffusioncontrolled and the magnitude of the diffusion current can
be expressed 2 2 ai = K(cs C e ) , where c, and Ce are the concentrations of the reactant in the bulk of the solution and
a t the electrode surface respectively, and K is a constant
whose value depends on conditions such as stirring rate and
the position and area of the electrode. At a sufficiently
high overpotential a limiting diffusion current is reached,
when ce = 0; the current is then proportional to G,, and i t
is this linear relation between current and iodine concentration which is utilised in the direct amperometric method.
A constant potential of 0.5 v was applied between an
inch-long platinum wire cathode and a Ag/Ag+ comparison
electrode, using a circuit similar to that shown by Charlot
31- took place a t
et ~ 1 . 2 2 ~
The reaction I,2ethe platinum electrode. The Ag/Ag+ half-cell was contained in a J-tube with a capillary outlet closed with agar
gel. It was fitted a t the upper end with a ground-glass
cone for insertion in the reaction flask, which was a closed
vessel of 500 ml. capacity containing 200 ml. of solution
stirred by a glass link stirrer a t 500 r.p.m. The change in
current with changing iodine concentration was followed
by measuring the potential drop across a 22 kR resistor
la J. 0. Edwards, Chew. Rev., 1952, 5
!, 455.

Treatise on Analytical Chemistry, ed. I. M. Kolthoff and


P. J. Elving, Interscience, New York, 1961, part 11, vol. 7 ; J. T.
Stock, Analyt. Chem., 1964, 36, 3 5 5 ~ .
2o E. C. Potter, J . A p p l . Chew., 1957, 7 , 285.
21 A. F. M. Barton, Ph.D. Thesis, University of Auckland,
1965.
19

2.5

5.0

7.5

1
0

lo6[ I 21 (MI
FIGURE
1 A direct amperometric kinetic plot
AB, Slope determination for reaction rate ; C, thiosulphate
addition; D, iodine addition ; DE, calibration determination
of reaction followed was generally less than 1% so that the
initial concentrations were practically unchanged. The
combined errors in solution composition and rate measurement resulted in z1 having a reproducibility of about f5%.
Solutions.-Analytical
grade potassium bromate and
potassium iodide were used, the solutions being standardised
with potassium iodate. Glacial acetic acid was distilled
from potassium dichromate and sulphuric acid ; the chloroacetic and P-chloropropionic acids were laboratory reagent
grade. The hydrogen ion concentration of each buffer
solution was obtained by means of a computer-calculated
successive approximation method which corrected for the
dimerisation of the carboxylic acids.24 Table 5 gives the
pK, values of the acids used at an ionic strength of 1 . 0 0 ~ .
Sodium perchlorate of suitable quality was not available
commercially, so i t was prepared from analytical grade
perchloric acid and anhydrous sodium carbonate by a
method incorporating precipitation and filtration of
22 G. Charlot, J. Fdoz-Lambling, and B. TrCmillon, ' Electrochemical Reactions, Elsevier, Amsterdam, 1962, ( a ) p. 22; (b)
p. 134.
z3 M. Spiro and A. B. Ravno, J . Chem. Soc., 1965, 78.
J. D. E. Carson and F. J. C. Rossotti, in * Advances in the
Chemistry of the Co-ordination Compounds,' ed. S. Kirschner,
Macmillan, New York, 1961, p. 180; N. G. Zarakhani and M. I.
Vinnik, Zhur. j i z . Khim., 1964, 38, 632.

View Article Online

1749

Inorg. Phys. Theor.


siliceous impurities. The sodium perchlorate stock solution was analysed by using the relation which was found to
hold between the molarity (M) and density (d) of sodium
perchlorate solutions in the range 2 - 6 ~ a t 25",

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

= 13.686d

- 13.794

Water for solution preparation was passed through an ionexchange purifier, then distilled from alkaline permanganate
with the use of a fractionating column to prevent carrythrough of spray. All solutions were made up to an ionic
strength I = 1 . 0 0 ~by adding sodium perchlorate. The
temperature was 25.00 & 0.05'.
Side Reactions.-The
concentrations of bromate and
hydrogen ion were such that association into bromic acid
was negligible.25 The tri-iodide concentration being
measured was so low that the fraction of iodide removed
in this form was very small. A t the hydrogen ion, iodine,
and iodide concentrations used, iodine hydrolysis and subsequent hypoiodite dispropotionation 26 were unimportant.
The rate of iodide autoxidation for the present conditions
calculated from the data of Herbo and Sigalla 27 was
negligible. The rate of the reaction
C1CH2C02H I- +ICH,CO,H
C1- calculated from
Wagner's rate constant 28 was insignificant. Hinton and
Johnston 29 found that molecular iodine was produced from
chloroacetic acid and iodide under conditions of greater
acidity and iodide concentration than those in the present
investigation ; however, the stoicheiometry of the related
iodate-iodide reaction in the presence of chloroacetic acid
has been
and in the present investigation there
was no observable variation in the iodide concentration
dependence. Hydrolysis of chloroacetic acid was found by
Hinton and Johnston 3 1 ~ 2 9to be negligible under conditions
similar to those in this work. In order to determine the
rate of loss of iodine by volatilisation, small quantities of
iodine were added to solutions similar to those in the kinetic
runs, and the iodine concentration was followed amperometrically. This effect proved to be small, and the greatest
correction necessary to the rate of any run was 3%.

expressed [B-][H+I2 instead of [HB][H+], but these alternatives are kinetically indistinguishable. Over a range of
buffer ratios only one buffer term appears in the rate law,
and hence only one form of the buffer, probably B-, is
catalytically active. The data are presented in Tables
2-4, and the rate constants, calculated by a least-meanTABLE1
Kinetic results in perchloric acid solution (concentrations
in M ; v in ?r~sec.-l; I = 1 . 0 0 ~ ;25')
109v

RESULTS

The rate law (1) found in previous bromate-iodide


investigations was confirmed in unbuffered perchloric acid
solutions (Table 1). The average value of KO from 9 runs
was used to calculate 1~ in Table 1, and there is good agreement with experiment over a fair range of reactant concentrations. Comparison of k , with previous work is
shown in Table 5.
The reactions carried out in carboxylate buffers showed
a marked enhancement of rate. Even a t the moderate
acetate concentration of O - O ~ M the
,
classical term found in
unbuffered solutions contributed a negligible amount to the
rate. The rate law for the three carboxylates investigated
proved to be

103[H+]
102[1-] 102[Br03-]
8.6
5.86
0-373
0.126
1.77
2.9
0.373
0.503
7.03
0.126
12.4
0.373
1.77
0.503
7.2
1-03
5.86
0-0759
0.250
3.2
3.05
0-708
0.250
8.1
7.03
0.0498
0,553
6.9
3.64
0.0758
0.552
2-8
1-77
0.0498
4.89
3.4
Calculated from rate law (1)with mean k , = 49

7.9
2.9
11.4
8.0
3.2
8.1
6.7
2.7
3.8
M - sec.-l.
~

TABLE2
Kinetic results in acetate buffers (concentrations in
M ; ZI in M sec.-l; I = 1 . 0 0 ~ ;25")
105[H+] 102[1-] 102[Br0,-] 1O2iCH3C0,-]

lO*[CH,CO,H]

100v
r
obs.
ca1c.a

11.3
0.95 25.0
2.38
10.9
40
40
11.3
3-78
0.503
2.38
10.9
2.9
3.2
11.3
6.19
0.250
2.38
10.9
2.2
2.6
2.16 6-19
4.89
9.84
8.6
8.1
7.5
11.5
1-89
4-89
0.504
2.37
3.9
3.9
11.5
3.10
4-89
0.504
2-37
6.3
6.4
2.29 6.19
4.89
2.03
1.89
1.94
1.81
2.16 3.10
4.89
9.8
8-6
4.1
3.7
2-22 6.18
4.89
6.00
5.42
5.0
4.9
2.19
6.18
4-89
7.94
7.07
6.1
6.2
2.19 6-18
2.5
1-95
7-94
7.07
2.7
2.26 6-18
4.89
4.03
3.71
3.3
3.4
11.5
3.09
1.95
1.46
6.82
6.9
6.6
11.4
3.09
1.95
1.92
8.9
8.4
8.4
11.6
3.09
1.95
0.99
4-65
4.9
4.7
11.6
3.09
11.7
4-65
11.3
4.89
0.99
5.97
1-03
4-89
4-86
11.8
4-9
4.8
6.29 6.19
4.89
0.515
1.32
4-3
3.9
6-13 3.09
2.74
6.85
8.2
8.6
4.89
6.06
1.03
4.89
3.81
9.4
3.9
3.9
6.24 3.09
4-89
1-64
4.17
5-1
5.5
5.97 3.10
4-86
11.8
1.96
6.2
5.7
a Calculated from rate law (2) with k o = 50 &r3
sec.-l and
kb = 5.4 x lo4 M - ~sec.-l.

squares computer program, appear in Table 5. When both


sides of rate law (2) are divided by the concentration
product, we obtain the expression
v/([H'l2[Br03-I[I-]) = KO

+ kbIB-1

(3)

Figure 2 presents the data of Tables 1-4 graphically by


z, = K,[H+]2[Br03-][I-]
kb[B-][H+]2[Br0,-][I-] (2)
means of equation (3), the intercepts and slopes yielding
where B- = CH3C0,-, C1CH2CH2C02-,and CICH2C02-. k , and kb respectively. The value of k , found by a leastFor convenience the acid dependence of the second term is mean-squares treatment of the data for each buffer agrees

25 J. Bjerrum, G. Schwarzenbach, and L. G. Sillen, ' Stability


Constants,' Chem. SOC.Special Publ. No. 7, 1958, pt. 11, p. 118.
26 K. J. Morgan, Quart. Rev., 1954, 8, 123.
27 J. Sigalla and C. Herbo, J . Chim. phys., 1957, 54, 733; C.
Herbo and J. Sigalla, Analyt. Chim. Acta, 1957, 17, 199.
C . Wagner, 2.phys. Chem., 1925, 115,121.

29

J. F. Hinton and F. J. Johnston, J . Phys. Chem., 1965, 69,

854.
30 S. N. Prasad and B. P. Gyani, J . I n d i a n Chem. Soc., 1954,
31,561.
31 J. F. Hinton and F. J. Johnston, J . Phys. Chem., 1963, 67,

2557.

View Article Online

1750

J. Chem. SOC.(A), 1968


TABLE3

Kinetic results in P-chloropropionate buffers


(concentrations in M ; er in M sec.-l; I = 1.00~;25')

Bronsted catalysis law 38a relating the catalytic constants


K b with the conjugate acid dissociation constants K,,

10%

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

10'[H+] 102[1-] 102[Br0,-] 102[C1CHsCH2C02-]102[C1CH2CH2C02H]obs.

1-14
1.16
1-15
1.16
1.17
1.16
1-17
4.39
4-28
4.40
4.38
4.28
0.340

0.373 4-89
7-89
7.69
6.19
4-89
0.82
0.81
1.03
4-89
6.17
6.06
0.373 25-0
4.41
4.37
3.09
1.95
2-63
2-62
3.09
1.95
4.41
4.37
3.09
4-89
2.63
2.62
1.03
0.553
3.87
14.5
1.03
0-98
0-84
3.07
1-03 0.98
2.39
9.0
0.372 1-95
1.63
6.09
0.709 0.98
0.84
3.07
3.10
4-89
14.8
4.33
Calculated from rate law (8) with k , = 42
k b = 1.7 x lo4 M - sec.-l.
~
Q

ca1c.a

3.3
7.0

3.3
7.4
7.5 7-3
11.6 10.0
4.1 4.0
6-5 6.5
9.6 10.1
7.6 7-7
3.5 3.4
8.6 8.8
4.4 4.4
2.5 2.4
4.3 4.5
- ~sec.-l and

TABLE4
Kinetic results in chloroacetate buffers (concentrations
in M ; TI in M sec.-f; I = 1*00~;
25")

10%

103[H+] 102[1-] lO*[BrO,-] 102[C1CH&Oa-] 1OZ[ClCHzCO2H] obs.

2.00 0.472 0.553


10.0
9.1
1-94 0.472 0.553
3-20
2-83
1-33 0.472 0.553
0.641
0.389
5.17
4-72
2-00 0.472 0.250
1-63 0.709 0.250
1-17
0.87
2-00 0.709 0.126
10.0
9.1
1.85 0.95 0.126
2.20
1.85
4.72
2-00 0.95
0.126
5.17
0.762 0.709 0.250
15.0
6-21
0.702 4.93
0.250
1.59
0.510
1.62
0.185 4.93
0.250
19.2
2-97
0.780 0.944 0.502
8.3
0-186 4.93
0.503
2.91
0.247
0.96
0.189 4.93 0.502
11.1
13-1
5.41 0.124 0.126
5.32
11.7
4.12
0.772 0.709 0.502
0.83
1-21
3.22 0.124 0.126
3-20
7.43
5-08 0.373 0.126
4-28
10.4
5.29 0.124 0.126
a Calculated from rate law (2) with K O = 50
k b = 2.5 X 10' M-' SeC.-l.

ca1c.a

32
31
12.4 12.6
3.2
3.0
8.3
8.4
3.9
3.7
12.5 10.6
4.5
4.2
9.2
8.4
4.9
4.3
5.4
5.4
2.4
2.2
6.2
7.4
0.76 1.04
2.9
2.8
9.1
8.3
6.8
7.2
1.19 1.13
16.4 15.6
6.9
6.8
M - ~sec.-l and

FIGURE
2 Graph of equation (3) for all data showing catalytic
effect of carboxylate ions
0 Unbuffered (mean of 9 experiments) ; A, CH,CO,-;
B, ClCH,CH,CO,-; C, ClCH,CO,DISCUSSION

General Mechanistic Considerations.-The rate laws


for the bromate-iodide reaction, bromate oxygen
exchange,53s34and in fact most halate-halide reactions,ls
exhibit second-order hydrogen ion dependence. Proton
catalysis can be considered to be due to the labilization
of the oxide ion separating from the halogen(v) atom.3k
The isotope effects in the bromate-iodide reaction,l3

TABLE5
Rate constants for rate laws (1) and (2) (25" except where indicated; k , in M-3 sec.-l and k b in
10-4kb
I
k0
BPKa
1.00
49 41
1.00
50 & 1
CH3*C024-61
5.4 f 0.1
ClCH2*CH2*C0,- 3-93
1.70 f 0.05
1.00
42 & 5
0.25 & 0.01
50 f 3
ClCH,*CO,3-66
17.8 (0')
82 a
110 a (30')
97
82 b
200 b
170
Activation energy 9.9 kcal. mole-'. b Isotope effect, k ~ l =
k 2.4.
~

1.00
Dilute solution ........................
Dilute solution ........................
Dilute solution ........................
Dilute solution ........................
Dilute solution ........................
D,O ....................................
Infinite dilution .....................
6

well with that in unbuffered solution. The rate constants


for the three carboxylates were found to obey closely the
32 R. P. Bell, ' Acid-Base Catalysis,' Oxford University Press,
London, 1941, (a)p. 82; (b) p. 143.
33 T. C. Hoering, R. C. Butler, and H. 0. McDonald, J . Amer.
Chem. SOC.,1956, 78, 4829.
94 M. Anbar and S. Guttmann, J . Amer. Chem. SOC.,1961, 83,
4741.

M-*

sec.-l; I in

bi)

Ref.
Table 1
Table 2
Table 3
Table 4
10
10
10
12
13
13
16

= 2.4, and bromate oxygen exchange,s3 kn/& =


1.7, imply equilibrium protonation prior to the ratedetermining step*32b Hinshelwood 36 has suggested in
the case of the bromate-bromide reaction that the rateKD/KH

35 J. 0. Edwards, ' Inorganic Reaction Mechanisms,' Benjamin,


New York, 1964, (a)p. 137; (6) p. 38.
36 C. N. Hinshelwood, J , Chem. SOC.,1947, 694.

View Article Online

1751

Inorg. Phys. Theor.


determining step may be the bimolecular reaction between the conjugate acids of the two anions shown in
reaction (4).

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

HBr

+ HBrO, 4 HBrO + HBrO,

(4)

This mechanism is attractively simple but it is unlikely


that two such strong acids would remain protonated long
enough to undergo a bimolecular collision.37 More
probably the necessary proton transfers will occur
simultaneously with the formation of an encounter between the reactants.
The intermediate Br02+has been postulated by several
workers,16*1sy38and rate law (1) is consistent with its
rapid formation and slow reaction with iodide. Nevertheless, we believe its existence is improbable for three
reasons. (a) Hoering et al.33 compared the rates of
bromate-water oxygen exchange and the bromate-halide
reactions. The oxygen exchange rate is given by
equation (5).
21

= 8.0

x 10-5[H+]2[Br0,-][H,0]

sec.-l a t 25" ( 5 )

If BrO,+ was the intermediate, then this must represent


the maximum rate possible for the bromate-halide
reactions. But these are all faster than the oxygen
exchange, even the bromate-chloride reaction,14 which
is the slowest of this series, as shown by equation (6).
"J

= 6-5 x

there must be either no change in the co-ordination


number, for example by Br-1-0 bond formation, or
else an increase with a Br-I bond. It should be noted
that, although the number of atoms round the central
atom remains unchanged when Br-0-1 bonds are formed,
the number of 0-orbitals associated with the central
bromine is increased to five, including two non-bonding
electron pairs, as in structure (11). The x-orbitals of the
double bonds are of lesser importance in determining
the stereochemistry and will be neglected in this discussion.41 In the reaction schemes shown, unshared electron pairs on the central bromine(v) are indicated by
bold lines in order to emphasize the stereochemistry.
The bromate ion has a pyramidal shape; 42 this can be
considered as a tetrahedral arrangement of electron pairs
around the central bronine atom. The molecule bromyl
fluoride, BrO,F, has been
and this can be
regarded as a model for the intermediates (111) and (V)
which appear in the reaction schemes. Intermediates
such as (VII) and (VIII) will have structures of the
trigonal bipyramidal type similar to a number of known
five-co-ordinate molecules.44
Reaction Mechanism iuz the Absence of Carboxylate
Catalysis.-Proposed
reaction paths are shown in
Scheme 1, which includes steps suggested by earlier

10-3[H+]2[Br03-][C1-] M sec.-l at 25" (6)

(b) The reaction between the oppositely charged ions


BrO,+ and I- would be fairly fast and under some
conditions the formation of BrO,+ by fission of the Br-0
bond would be expected to be rate-limiting, giving a rate
law z1 = K[H+]2[Br0,-]. This rate law has never been
observed for any bromate reaction (except oxygen exchange). Hence bromate reactions differ from some
reactions of nitric acid 39 which have a rate law z1 =
k[H+]2[N0,-], indicating the existence of NO,+ as an
intermediate. (c) The results in Figure 2 show that
carboxylate ions are effective catalysts, but it is dificult
to imagine why this should be so if BrO,+ was the
intermediate. Indeed, the occurrence of catalysis by
bases (nucleophiles)is probably a good indication that no
cationic intermediate such as BrO,+ is formed. In the
analogous case of nitrous acid reactions, Hughes, Ingold,
and Ridd 40 have found catalysis by a variety of nucleophiles, and they were able to prove that free NO-t was
not an intermediate.
In an oxyanion reaction it is necessary to consider
whether the co-ordination number of the central atom
will increase, decrease, or suffer no change in the transition states and intermediates that are formed. The
evidence given above shows that there is no decrease in
co-ordination number to form Br02+. This means that

SCHEME

w o r k e r ~ . l ~ $ The
~ ~ * first
~ * step is the substitution of
iodide either directly on the central bromine atom or by
formation of an 0-1 bond, and probable transition states
for this step are shown. It is also possible that the
structures (I) and (11) may exist as true intermediates
rather than transition states. The first step is followed
by nucleophilic attack of water or iodide on the iodine
R. J. Gillespie and R. S. Nyholm, Quart. Rev., 1957, 11, 339.
J. W. Mellor, ' Comprehensive Treatise on Inorganic and
Theoretical Chemistry,' suppl. 11, part I, Longmans, London,
1956, p. 798.
43 M. Schmeisser and E. Pammer, Avzgew. Chem., 1957, 69, 781.
44 E. L. Muetterties and R. A. Schunn, Quart. Rev.,1966, 20,
245.
41

37

H. Zimmermann and J. Rudolph, Angew. Chem. Internat.

Edn., 1965, 4, 40.


38 E. Abel, Helv. Chim. A d a , 1950, 33, 785.
39 C. A. Bunton and E. H. Halevi, J . Chem. Soc., 1952, 4917.
40 E. D. Hughes, C. K. Ingold, and J. H. Ridd, J . Chem. Soc.,
1958, 85.

42

View Article Online

J. Chem. SOC.(A), 1968

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

1752
atom. There has been mention in the literature34that
some induced bromate oxygen exchange with water occurs
during the bromate-iodide reaction. If correct, this
would imply that the first step is reversible, and therefore
the second step must be a rate-determining attack by a
water molecule, in order to agree with the observed rate
law (1). The primary products in Scheme 1 are assumed
to undergo further fast reactions with iodide giving
finally iodine and bromide. The halogen compounds
of these types which can be isolated react rapidly with
iodide; in other cases this may be inferred from indirect
observation^.^^

Br0,BrO-

+ 2H+ + 21- --+ I, + BrO- + H,O

+ 2H+ + 21- +I, + Br- + H,O

Anbar and Guttmann34 found that the chloridecatalysed oxygen exchange between bromate and water
obeyed rate law (7).

=5

x 10-1[H+]2[Br03-][C1-]M sec.-l at 25"

I
or

-1- H,0

J3r,

0 II

O-CI

CH,-C-0-Br=O

(7)

This has the same form as rate law (6) found for the
bromate-chloride redox reaction, but the oxygen exchange rate is 77 times greater. Anbar and Guttmann
deduced the existence of two different intermediates,
for example ClBrO, and OBrOC1, one for each reaction.
However a mechanism such as Scheme 2 with an equili-

&Br,
-{- H,O
(1') 0 II c1

ing step. It is preferable to describe this as nucleophilic


catalysis,35brather than base catalysis, since the atom
attacked is bromine(v) rather than hydrogen. Nucleophilic catalysis occurs in some other oxyanion reactions,
for example nitrosation by nitrite4* and hydrolysis of
dichromate,46 but has not been previously reported for
bromate reactions. I n a detailed treatment of the
mechanism it is necessary to consider the following
aspects of the reaction. (a) The order in which carboxylate and iodide attack bromate; on present evidence it is
not possible to decide which nucleophile attacks first.
(b) The point of initial attack by iodide; there is no
evidence to distinguish between attack on oxygen or
attack directly on the central bromine atom. Two
possible reaction paths are illustrated in Scheme 3.
(c) The point of attack by the carboxylate ion; if it
forms a bond to bromine we have a structure which is
a mixed acid anhydride, as shown for the case of acetate.

PI)

"It

II

0
There is also the possibility of attack on the previously
substituted iodine, removing it without its bonding
electrons to form iodonium acetate.

CH3C02- --t Br0,CH,CO,I


Br0,I
CH,CO,I
H,O ---w CH3C0,H
HOI

slow
fast

The observed acceleration in the presence of carboxylates


would then be explained by carboxylate attack on the
iodine atom a t a faster rate than the corresponding
reaction with water in Scheme 1.
If we assume attack by the carboxylate ion on the
central bromine atom, either as a preliminary reaction
or subsequent to a reaction between bromate and iodide,
Scheme 3 can be formulated. The intermediate, (VII)

/H2O

HBr0,

4- HOCI'

SCHEME
2

brium followed by a slower attack on the intermediate


by H,O would explain the rate difference without invoking different intermediates in the two reactions. The
00first step involves reversible removal and addition of
H,O on the central bromine atom, and the rate of this
step would equal the observed oxygen exchange rate.
I n the second step the water molecule attacks and
removes the chlorine atom giving the initial products
of the redox reaction. But a t this stage it is impossible
to know which of the alternative intermediates, (V)
I, 4- Br0,- -+ B- HOI +- HBrO, + B- I, + Br0,- + J3or (VI), is actually present. It is impossible to determine
SCHEME 3
uniquely a mechanism from kinetic data only, and hence
some of the possible alternatives have been given in the or (VIII), with co-ordinated carboxylate, must exist in
proposed reaction schemes.
greater concentration or be more reactive than the
Reaction Mechanism with Carboxylate Catalysis.-The
corresponding intermediate, (111) or (IV), of Scheme 1, in
appearance of the carboxylate catalysis term in rate
45 H. A. Young, J . Amer. Chem. SOC.,1950, 72, 3310.
law (2) indicates that bromate, iodide, and carboxylate
46 B. Perlmutter-Hayman and M. A. Wolf, J . Phys. Chem.,
occur together in the transition state of the rate-determin- 1967, 71, 1416.

View Article Online

Inorg. Phys. Theor.

1753

Published on 01 January 1968. Downloaded by University of Central Florida on 31/10/2014 14:34:57.

order to give effective catalysis by carboxylates. To


conform with rate law (2) the first step may be slow,
followed by fast attack by water or another iodide;
or the first step may be a fast equilibrium followed by
rate-determining water attack on the iodine atom. In
TABLE6
Rate constants for reactions of bromate with X according
3
I in M ; activation energy
to rate law (8) ( K in ~ - sec.-l;
E , in kcal. mole-I; 25")
Ref.
47
8 b
48
HO,-a ......
I- ............
9.9
Table 1, 10
Br - ............
15.8
14, 48
c1- ............
14
13.6
33
H,O .........
NO,-'
......
49
s,o,2- ...... 18
a The observed rate laws were converted t o the form of
equation (8) using the dissociation constants of the conjugate
acids, [HX+] = [H+][X]/K,. 6 AH for dissociation of H,O,
was used to correct the observed activation energy t o the
value required by equation (8).

so,,-

X
a

......

2.10
0.06
1.00
0.26
1.18
0.90
0.01

9.7 x 109
2.4 x lo8
49
2.2
6.5 x
8.0 x 10-5
105

E a

his investigation of the reactions of bromate with


bromide and chloride, Sigalla 14 discovered a new term
in the rate law, k[H+]2[BrO-,] [Br-][Cl-], when bromide
and chloride were both present. In the case of these

weaker nucleophiles the second halide probably functions


in the same way as B- in Scheme 3.
Other Bromate Reactions.-A number of reactions of
bromate (Table 6) have been found to follow rate law
law (9,where X is the relevant nucleophile.
z, =

-d[BrO,-]/dt

= K[H+]2[Br0,-] [XI

Various mechanisms have previously been put forward


for these reactions,50 but it appears that they can be
all explained satisfactorily by Scheme 1with X replacing
I-. Thus, in the case of the bromate-sulphite r e a ~ t i o n , ~ l
isotopic experiments showed that the oxygen atom
gained by the sulphite ion came from bromate. Therefore, the sulphur atom of sulphite attacks an oxygen
atom of the bromate ion, as in structure (11). The
reactions of bromate are characterized by nucleophilic
attack by the reducing agent resulting in transfer of
electrons towards the central bromine(v) and a temporary increase in its co-ordination number.
The authors thank the New Zealand University Grants
Committee for a Research Fund Fellowship and an Internal
Post-Graduate Scholarship (to A. F. M. B.), and for an equipment grant.
[7/1659 Received, Decembev 18th, 19671
49

W. G . Lowe and D. J. Brown, 2. anorg. Chem., 1934, 221,

173.
F. A. Williamson and E. L. King, J . Amer. Chem. SOC.,1957,
79, 5397.
4 8 H. A. Young and W. C . Bray, J . Anzev. Chern. SOC.,1932,
54, 4284.
47

50 T. A. Turney, ' Oxidation Mechanisms,' Butterworths,


London, 1965.
51 J. Halperin and H. Taube, J . Amer. Chem. SOC.,1952, 74,
375.

Vous aimerez peut-être aussi