You are on page 1of 8

Ind. Eng. Chem. Res.

1993,32,144S1456

1449

Kinetics of Hydrogen Peroxide-Chlorate Reaction in the Formation of


Chlorine Dioxide
Michael Burke: Joel Tenney,t Bhart Indu: M. Fazlul Hoq: Shelley Carr,+and
W. R. Ernst**+
School of Chemical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332,and Eka Nobel,
Inc., Marietta, Georgia 30075

This paper reports and discusses an experimental program that investigates chlorine dioxide formation
by reduction of sodium chlorate with hydrogen peroxide in aqueous sulfuric acid. The rate of
reaction was studied in both batch and well-mixed reactors over ranges of temperature and reactant
concentrations which bracket conditions of commercial interest. The rate data were correlated by
a power law model in which an acidity function H-was used to characterize the acidity of sulfuric
acid. The acidity function takes into account the large influence on the acidity of sulfuric acid
exhibited by both byproduct sodium sulfate and sodium ions associated with sodium chlorate. The
rate model reasonably predicted transient chlorine dioxide production rate data gathered in a
commercial plant trial. The rate model should provide a useful tool for design.

Introduction
Chlorine dioxide is an oxidizing gas which, in aqueous
solution, is used in water purification and as a chemical
bleach. The largest volume of chlorine dioxide is consumed
by the paper industry, where, because of environmental
concerns, it is replacing molecular chlorine in the chlorination sequence of the pulp bleaching process. Unlike
chlorine, however, chlorine dioxide is relatively unstable
and cannot be economically transported or stored over
long periods without substantial loss by decomposition.
These problems are avoided by locating chlorine dioxide
plants at the pulp mill site and operating them so as to
satisfy the daily demand of the mill.
Chlorine dioxide can be produced by reacting sodium
chlorate in sulfuric acid solution with a reducing agent
such as sodium chloride, methanol, or hydrogen peroxide
(Tenney et al., 1990; Norell, 1988; Ernst et al., 1988;
Holmstrom et al., 1987; Isoa Isa et al., 1983; Masschelein,
1979;Rosen, 1976;Atkinson, 1974;Rapson, 1958; Sprauer,
1956; Soule, 1943). One widely used process involves
continuously feeding sodium chlorate, sulfuric acid, and
a reducing agent to a single vessel (generator) that serves
as a reactor, evaporator, and crystallizer (see Figure 1).
This type of process is known as the single-vessel process
(SVP). Product gases exit through a vapor pipe, flow
through a condenser, where the water vapor is condensed,
and pass into a packed column, where chlorine dioxide is
absorbed in water. Gaseous byproducts, either chlorine,
carbon dioxide, or oxygen, depending upon the choice of
reducing agent, vent from the top of the column and either
pass into a second absorber or vent to the atmosphere. A
slurry of sodium sulfate crystals is continuously pumped
from the bottom of the generator to a filter, where the
crystals are removed from the process. Filtrate is returned
to the generator.
The choice of reducing agent has been influenced by
environmental concerns over the byproducta. Until recently, most chlorine dioxide was produced by reducing
sodium chlorate with sodium chloride. This process has
two shortcomings: (1)it yields chlorine as a major gasphase byproduct; and (2) it yields substantial amounts of
solid sodium sulfate byproduct.

* Author to whom correspondence should be addressed.


+ Georgia Institute of
8 Eka Nobel, Inc.

Technology.
0888-5885/93/2632-1449$04.00/0

To eliminate problems associated with the chloridechlorate process, an increasing number of plants are using
methanol as the reducing agent. The methanol-chlorate
process produces no chlorine but, rather, oxidation products of methanol. Complete oxidation of methanol leads
to efficient utilization of methanol, described by the
stoichiometry
6NaC10,

+ CH,OH + 4HzS04

6C10z + COz +
5Hz0 + 2Na3H(S04), (1)

The obvious advantage of this process is that carbon


dioxide is relatively inert in the environment. In present
processes, carbon dioxide is vented to the atmosphere.
Although the methanol-chlorate process is a significant
improvement over the chloride-chlorate process, it is not
without shortcomings. One problem arises from the fact
that chlorate oxidizes methanol in a stepwise manner, first
forming formaldehyde,then formic acid, and fiially carbon
dioxide (Hoq et al., 1991b,c, 1992;Indu et al., 1991a).The
final step of this process, the oxidation of formic acid to
carbon dioxide, is a relatively slow step at the high acidic
conditions of a commercial generator. We recently found
significant concentrations of methanol and formic acid in
chlorine dioxide absorber solutions from several commercial chlorine dioxide plants, indicating that some of the
methanol fed to the processes and some of the intermediate, formic acid, do not react but, instead, volatilize and
escape from the generator along with the chlorine dioxide
(Hoq et al., 1991a). This problem most likely arises because
highly volatile methyl formate is produced in the commercial generators (Indu et al., 1992). These findings
demonstrate that both carbon dioxide and formic acid are
byproducts of this process. The process stoichiometry is
therefore not represented simply by eq 1but rather by a
combination of eq 1 and the incomplete oxidation of
methanol to formic acid:
6NaC10,

+ l.BCH,OH + 4H2S0,

6C102+
1.5HCOOH + 4.5H20 + 2Na3H(S04), (2)

A second problem is associated with the fact that the


methanol-chlorate process requires high sulfuric acid
concentrations to achieve the desired production rate of
chlorine dioxide in reasonably sized equipment. At the
high sulfuric acid concentrations, crystalline sodium
sesquisulfate, or acid salt cake, forms, as shown in eq 2.
0 1993 American Chemical Society

1450 Ind. Eng. Chem. Res., Vol. 32, No. 7, 1993

A m .

chll1.d
valor

Y1.C.
Vanb

Pump

Pump

Pump

Figure 1. Configuration of a commercial SVP chlorine dioxide generator.

In the past, most of the salt cake could be consumed in


kraft paper mills, which utilize it to make up for losses in
the pulping and alkaline chemical recovery processes.
Unconsumed salt cake was sent to waste after it was
neutralized with sodium hydroxide. In recent years, most
mills have made major equipment and operating modifications in the pulping and sulfur-recovery processes
which have eliminated much of the sulfur losses and closed
the sulfur material balance. Now, most of the salt cake
byproduct of the chlorine dioxide process must be neutralized and sent to waste. This has become an increasingly
costly process.
Some of the problems associated with the methanolchlorate process might be eliminated if hydrogen peroxide
were used in reducing chlorate to form chlorine dioxide.
The hydrogen peroxide-chlorate process follows the
stoichiometry
2NaC10,

+ H202+ H2S0,

2C10, + 0, + 2H20 +
Na2S0, (3)

This process was developed to address the shortcomings


of methanol- and chloride-based processes. At first glance,
the high cost of hydrogen peroxide would appear to rule
out the economic feasibility of the hydrogen peroxidebased process; however, individual mills must weigh this
cost against the followingprocess advantages and potential
cost reductions in light of their specific economic factors
when the decision of which technology to use is made: (1)
at similar temperatures, similar acid concentrations,
similar chlorate concentrations, and stoichiometric chlorate-reducing agent ratios, the hydrogen peroxide-chlorate
process generates chlorine dioxide at a faster rate than
the methanol-chlorate process; (2) as a consequence of
(l),the hydrogen peroxide-chlorate process can achieve
reasonably fast chlorine dioxide generation rates at lower
sulfuric acid concentrations where crystalline sodium

sulfate or neutral salt cake forms, as shown by eq 3; (3)


the hydrogen peroxide-chlorate process produces less salt
cake per mole of chlorine dioxide; (4) also as a consequence
of (11, smaller, less expensive reaction equipment might
be required for a given chlorine dioxide production rate;
(5)contamination of the chlorine dioxideproduct by formic
acid would be eliminated; and (6) byproduct oxygen might
be recoverable at sufficient purity for use within the plant.
Although the stoichiometry of the hydrogen peroxidechlorate reaction, eq 3, was established previously, no
kinetic data for generating chlorine dioxide by this process
have been reported in the open literature. The purpose
of this work is to study the kinetic behavior of this reaction
over a wide range of temperatures and reactant concentrations and to develop a kinetic model that can be used
in design studies of this process.

Experimental Section
a. Flow Reactor. The reactor (I in Figure 2) is a
cylindricalPyrex vessel, 20.6 cm high X 8.9 cm i.d., operated
with a 300-mL liquid volume. It has three openings at the
top which connect to a Teflon-coated thermocouple, Pyrex
heating coil, and gas outlet tubing. The thermocouple is
part of a temperature control system (C, E-H) which
maintains reactor liquid temperature to within f0.5 OC.
The gas outlet tubing connects to a Pyrex manifold,
equipped with vacuum gauge (M), Teflon-coated thermocouple (D), gas absorbers for sampling (0and P), and
an aspirator-needle valve assembly for maintaining vacuum (Qand N). All connections and fittings are made of
PFA or TFE Teflon, both of which remain flexible and
inert in a chlorine dioxide environment.
Peristaltic pumps A and B continuously feed 5 mL/min
aqueous sodium chlorate and 5 mL/min aqueous sulfuric
acid-hydrogen peroxide solution, respectively, through

Ind. Eng. Chem. Res., Vol. 32, No. 7, 1993 1451

r--------------------

.-------.
--------

I C-----------,

F I /

Q@

I
I
I
I
I

t l

openings in the side and bottom of the reactor. Liquid


effluent is continuously removed by a peristaltic pump
(R) to maintain a constant liquid level. Air is fed into the
bottom opening of the reactor along with the sulfuric acidperoxide solution at a rate of 500-1000 cm3/min. The air
is inert in this system and provides vigorous agitation,
strips chlorine dioxide from the reaction solution, and
dilutes chlorine dioxide in the vapor space of the reactor
and manifold. The reactor operates under a ca. 0.5-atm
vacuum. The combination of air dilution and vacuum
provides a means of maintaining chlorine dioxide partial
pressure below about 0.1 atm, above which chlorine dioxide
rapidly decomposes with a sudden pressure increase.
Reaction gases and dilution air flow from the reactor to
the manifold, which is maintained at a sufficient temperature by heating tape to prevent water from condensing.
The gaseous products and dilution air pass through the
manifold and through two gas absorbers (J and K)
containing sodium sulfite solution,which removes chlorine
dioxide by reducing it to chloride. The air exiting the gas
absorbers passes out of the system through the aspirator
(N).The manifold is wrapped with black tape to prevent
decomposition of chlorine dioxide by UV light.
Liquid and gaseous effluents are sampled after the
system has been operated for four liquid residence times.
The liquid effluent is sampled by diverting a measured
flow from pump R into a large measured volume of cold
water to quench the reaction. For gas sampling, the gas
absorbers, J and K, are replaced with two other absorbers;
each contain 500 mL of 10wt % potassium iodide solution
buffered with sodium phosphate. The gas is sampled by
diverting the flow through these absorbers for a measured
period of time.
Mixing was characterized in the reactor under nonreactive conditions by un-steady-state tracer experiments
(Smith, 1981). The tracer, sodium carbonate, was monitored in the effluent liquid after its concentration was
increased stepwise in the feed. The reactor was initially
filled with 300 mL of water and agitated by flowing air
into the bottom. At time t = 0, the two feed pumps were
started, one feeding water and the other feeding a sodium
carbonate solution of concentration Co. Simultaneously,

the effluent pump was started up at a rate equal to the


total input rate. The effluent was sampled periodically
and analyzed for tracer. Several system parameters were
varied to determine whether they influenced mixing.
During reaction experiments,gas samples were analyzed
for chlorine dioxide and chlorine by titrating the contents
of the absorbers at neutral and acidic conditions with 0.1
N sodium thiosulfate using a single pin Pt electrode and
a Ag/ AgCl reference electrode (no chlorine was detected
in this study). The effluent liquid was analyzed for
dissolved chlorine dioxide by measuring the absorbance
of UV light at 370 nm and for unconverted reactants by
various titration procedures: hydrogen peroxide was
determined by first bubbling nitrogen through the reactor
liquid to strip out dissolved chlorine dioxide, akidifying
with sulfuric acid, and titrating with 0.1 N potassium
permanganate; sulfuric acid was determined by titrating
with 0.1 N sodium hydroxide; and chlorate was determined
according to a technique reported by Aieta et al. (1984)
and Vogel (1978). All analyses were triplicated using a
Fisher Scientific computer-aided titration apparatus. A
deviation in results of less than 2 % was considered to be
satisfactory.
b. Batch Reactor. The equipment consists of a UV
spectrophotometer (Milton Roy Spectronic 1201) with a
0.4-cm3flow cell of 1-cm path length. The reaction vessel
is an open, 125-mL stirred flask, partially submerged in
a constant-temperature bath ( i 0 . 5 "C). Total volume of
the reaction solution was 100 mL in each experiment.
Liquid was continuously pumped at 100mL/min from the
flask through the flow celland back to the flask. Residence
time in this recycle system was about 6 s.
An aqueous solution of sodium chlorate and sulfuric
acid was placed in the flask, rapidly agitated with a
magnetic stirrer, and circulated through the UV flow cell.
The reaction was begun by adding the appropriate volume
of hydrogen peroxide to the flask and monitoring the
absorbance at 370 nm as a function of time. The
concentration of chlorine dioxide was determined from
the absorbance on the basis of an extinction coefficient,
370 = 1040 cm-l M-l.

1452 Ind. Eng. Chem.Res., Vol. 32, No. 7,1993


Table I. Exwrimental Flow Reactor Data
reactor solution concn, M
expt
temp,OC
H202
NaClOs
HBO,
Nap904
a
b
Runs without Added Sodium Sulfate
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

55
55
55
55
55
55
55
56
56
57
57
57
57
55
55
55
55
55
55
54
54
44
45
55
55
44
44

0.63
0.49
0.48
0.73
0.73
0.50
0.50
0.13
0.13
0.13
0.14
0.13
0.12
0.16
0.15
0.08
0.07
0.13
0.13
0.12
0.12
0.15
0.15

0.43
1.57
2.69
0.76
0.95
1.82
1.21
3.03
1.07
1.72
1.72
2.63
2.48
2.67
2.73
3.06
2.98
2.98
3.09
3.10
3.16
1.34
1.32
1.48
1.56
1.42
1.43
0.77
0.88
1.44
1.43
2.43
2.38
2.84
2.82
2.85
3.07
2.05
2.20
2.17
2.20
2.20
2.18
2.45
2.52
2.55
2.51
2.48
2.08
1.98
1.95
2.02
0.83
0.93
0.98
0.97
0.48
0.45
0.49
0.51

60
60
60
60
60
62
59
61
63
60
58
60
60
60
65

0.39
0.20
0.40
0.39
0.23
0.20
0.29
0.28
0.48
0.46
0.46
0.51
0.52
0.50
0.49

2.13
1.67
2.03
1.82
1.64
1.57
2.14
2.10
1.67
1.58
1.46
1.44
1.02
1.72
1.32

50
40
50
50
40
40
50
40
40
45
45
50
70
60
65
45
45
55
65
85
85
58
58
60
60
60
60
53
53
40
40
70

70

0.83
2.22
0.67
0.31
0.46
0.92
0.54
0.84
0.58
1.33
1.31
0.67
0.12
0.19
0.053
0.24
0.034
0.084
0.029
0.001
0.0005
0.41
0.39
0.49
0.51
0.50
0.50
1.38
1.41
2.81
2.79
1.53
1.65
0.12
0.12
0.49

0.50

3.90
2.00
1.90
3.56
3.64
3.36
3.30
2.04
3.56
2.68
2.60
1.84
1.80
1.88
1.92
1.04
2.02
1.52
1.04
2.16
2.14
1.92
1.98
1.00
1.06
1.01
1.00
3.63
3.61
2.09
2.05
1.66
1.67
1.46
1.49
1.25
1.28
0.95
1.27
1.24
0.75
0.75
0.81
0.80
1.22
1.24
0.75
0.77
1.30
1.26
0.78
0.78
3.63
3.65
3.65
3.68
3.31
3.35
3.28
3.26
1.85
2.70
2.00
2.03
2.81
2.88
2.14
2.03
2.04
2.03
2.06
1.16
1.56
1.62
1.54

0.10
0.00
0.10
0.44
0.36
0.64
0.70
0.00
0.44
0.32
0.40
0.16
0.20
0.12
0.00
0.00
0.00
0.00
0.46
0.00
0.00
0.08
0.02
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.34
0.33
0.00
0.00
0.00
0.00
0.04
0.00
0.01
0.00
0.00
0.00
0.00
0.03
0.01
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.09
0.05
0.12
0.14

0.98
1.00
0.95
0.89
0.91
0.84
0.83
1.00
0.89
0.89
0.87
0.92
0.90
0.94
1.00
1.00
1.00
1.00
0.69
1.00
1.00
0.96
0.99
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
0.83
0.84
1.00
1.00
1.00
1.00
0.96
1.00
0.99
1.00
1.00
1.00
1.00
0.98
0.99
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
0.97
0.99
0.96
0.96

4.00
2.00
2.00
4.00
4.00
4.00
4.00
2.04
4.00
3.00
3.00
2.00
2.00
2.00
1.92
1.04
2.02
1.52
1.50
2.16
2.14
2.00
2.00
1.00
1.06
1.01
1.00
3.63
3.61
2.09
2.05
2.00
2.00
1.46
1.49
1.25
1.28
0.99
1.27
1.25
0.75
0.75
0.81
0.80
1.25
1.25
0.75
0.77
1.30
1.26
0.78
0.78
3.63
3.65
3.65
3.68
3.40
3.40
3.40
3.40

-H-uncorr
for NaClOs

-H-corr
for NaClOs

1.49
0.41
0.35
1.30
1.34
1.18
1.14
0.44
1.30
0.80
0.75
0.31
0.28
0.34
0.37
-0.17
0.43
0.12
-0.30
0.51
0.50
0.36
0.40
-0.20
-0.16
-0.19
-0.20
1.34
1.33
0.47
0.44
0.18
0.18
0.09
0.11
-0.04
-0.02
-0.24
-0.03
-0.05
-0.35
-0.35
-0.32
-0.32
-0.06
-0.05
-0.35
-0.34
-0.01
-0.04
-0.34
-0.34
1.34
1.35
1.35
1.37
1.17
1.19
1.15
1.14

1.54
0.60
0.67
1.39
1.46
1.40
1.29
0.80
1.43
1.01
0.96
0.63
0.58
0.66
0.69
0.20
0.78
0.48
0.07
0.88
0.88
0.52
0.56
-0.02
0.03
-0.02
-0.03
1.43
1.44
0.64
0.62
0.47
0.47
0.43
0.44
0.30
0.35
0.01
0.24
0.21
-0.09
-0.09
-0.06
-0.03
0.24
0.26
-0.05
-0.04
0.24
0.20
-0.10
-0.09
1.44
1.46
1.47
1.48
1.22
1.24
1.21
1.20

1430
161
900
1860
1480
3220
2020
469
1330
1590
1600
958
1190
830
672
30
73
143
54
721
280
429
478
47
59
33
29
5347
5047
169
208
6125
5521
132
141
182
225
25
92
98
15
13
32
34
66
85
19
18
38
47
10
10
1356
1376
852
851
438
468
240
220

0.19
0.74
0.32
0.34
0.75
0.80
0.35
0.26
0.24
0.23
0.21
-0.46
-0.15
-0.12
-0.08

0.45
0.94
0.56
0.56
0.95
0.99
0.60
0.61
0.44
0.42
0.39
-0.29
-0.03
0.01
0.08

617
1420
657
842
1620
2260
229
373
156
128
73
4
10
18
38

exper R ~ o , ,
1V M/min

Runs with Added Sodium Sulfate


61
62
63

64
65
66
67
68
69
70
71
72
73
74
75

1.15
1.30
1.00
0.97
1.94
1.87
1.61
1.72
1.96
1.97
2.44
2.00
2.37
2.47
1.47

0.62
0.68
0.67
0.68
0.59
0.61
0.57

0.54
0.51
0.51
0.46
0.37

0.40
0.40
0.51

3.00
4.00
3.00
3.00
4.75
4.75
3.75
3.75
4.00
4.00
4.50
3.16
3.93
4.09
3.01

Ind. Eng. Chem. Res., Vol. 32, No. 7,1993 1453


Table I. (Continued)
expt

ternp,OC

HnO2

76
77
78
79
80
81
82
83

65
65
67
68
65
54
54
55

0.46
0.49
0.45
0.44
0.42
0.11
0.10
0.10
0.09
0.08
0.08
0.29
0.28
0.03
0.07
0.07
0.10
0.10
0.09
0.09
0.09
0.10

84

54

85
86
87

65
65
65
65
55
55
55
47
46
57
57
45
44

88

89
90
91
92
93
94
95
96
97

reactor solution concn, M


NaClOa
H&h
Na2SO4
a
b
Runs with Added Sodium Sulfate
1.39
1.57
1.44
0.52
3.01
1.42
1.63
2.11
0.44
3.74
1.31
2.04
1.47
0.58
3.51
1.28
2.04
1.47
0.58
3.51
1.32
2.14
2.06
0.49
4.20
2.54
2.01
1.44
0.58
3.45
2.57
2.06
1.39
0.60
3.45
2.29
2.52
1.46
0.63
3.98
2.45
2.59
1.39
0.65
3.98
2.60
2.08
1.42
0.59
3.50
2.65
2.08
1.42
0.59
3.50
2.20
2.28
1.62
0.58
3.90
2.11
2.35
1.55
0.60
3.90
2.61
2.31
1.47
0.61
3.78
1.02
3.04
1.00
0.75
4.04
1.07
3.04
1.00
0.75
4.04
1.10
3.03
1.00
0.75
4.03
1.08
3.08
1.00
0.75
4.08
0.53
2.99
1.00
0.75
3.99
0.51
3.05
1.00
0.75
4.05
0.53
3.00
1.00
0.75
4.00
0.56
3.04
1.00
0.75
4.04

Results
a. SystemCharacterization. We ran the flow reactor
tracer experimentsunder nonreaction conditions primarily
to establish the rate of air flow that would provide enough
mixing so that the reactor could be modeled as a constant
flow stirred tank reactor (CSTR). We also examined how
other parameters would influence mixing, such as liquid
volume in the reactor, feed flow rate, pressure in the
reactor, and feed location of the tracer solution. Figure
3 shows tracer response data gathered in two experiments,
one in which air flow was 108 cm3/min and one in which
there was no air flow. Other conditions were identical in
the two experiments: liquid volume was 300 mL, total
liquid rate was 20mL/min, pressure was atmospheric, and
tracer was fed into the bottom of the reactor. Figure 3
shows that, with no air flowing, the reactor approached
that of a plug flow reactor (PFR), as one might expect.
With air flowing, the data points follow closely to the solid
line, which was calculated by the transient CSTR equation

C/C,= 1- exp(-t/r)

(4)
where the value of T was set equal to the experimental
space time, 15 min.
When we varied the other parameters of the system or
fed the tracer into the side of the reactor instead of the
bottom, the reactor continued to behave as a CSTR as
long as the air flow was above about 100 cm3/min. These
results demonstrate that we could assume CSTR behavior
throughout our reaction experiments since we always
maintained the air rate well above 100cm3/minto provide
sufficient dilution of the chlorine dioxide.
We reacted several solutions containing hydrogen peroxide (as limiting reagent), sodium chlorate, and sulfuric
acid, both in substantial excess, to complete conversion of
hydrogen peroxide to check the stoichiometricCoefficients
of the reaction producing chlorine dioxide. The reaction
was conducted in a batch reactor. Nitrogen was bubbled
through the reaction liquid for agitation and for sweeping
product gases to an absorber containing a solution of
sodium sulfiteand sodium bisulfite. This solutionconverts
chlorine dioxide to chloride. This process was continued
until the reactor solution became colorless. An analysis
of the absorber solution for chloride, and of the reactor
solution for unreacted chlorate and acid, yielded the
following stoichiometric coefficients based on 1 mol of

-H-uncorr
for NaClOa

-H-corr
for NaClOs

-0.05
-0.08
0.29
0.29
0.24
0.27
0.31
0.61
0.66
0.32
0.32
0.44
0.49
0.48
0.97
0.97
0.97
0.99
0.94
0.98
0.95
0.97

0.11
0.09
0.45
0.45
0.40
0.58
0.62
0.89
0.96
0.64
0.64
0.70
0.75
0.79
1.09
1.10
1.10
1.12
1.01
1.04
1.01
1.04

exper RCQ,
1V M/min

36
32
208
347
128
75
82
326
466
258
249
1390
1300
99
420
408
211
225
138
153
59
64

chlorine dioxide formed hydrogen peroxide,0.52;sodium


chlorate, 0.99;and sulfuric acid, 0.51. These coefficients
agree quite well with those in eq 3. The slight disagreement
between measured and theoretical values may have
resulted from two possible sources: error in determining
chlorate and acid consumption, since both were present
in excess, and slight decomposition of hydrogen peroxide.
b. Kinetic Study. For the kinetic study, we used both
batch and flow reactors so that we could measure a wide
range of rates which spanned several orders of magnitude.
For gathering rates at the lower end of the scale, we found
the batch system was more reproducible; at high rates the
on-line analytical equipment could not respond quickly
enough to monitor accurately the change in chlorine
dioxide with time. At the middle and higher end of the
rate range, the flow reactor was more accurate and provided
good reproducibility. The flow reactor was limited at the
higher end of the rate range because of foaming.
It is evident from eq 4 that steady state can be achieved
in the flow reactor after four residence times under
nonreactive conditions. A simple calculation shows that
under reaction conditions steady state can be achieved
more quickly depending on the reaction rate. In any case,
in this study all flow reactor samples were taken after four
residence times.
We gathered the flow reactor data over the temperature
range 40-85 "C and feed concentrations 0.05-2.9 M
hydrogen peroxide, 0.5-3.2 M sodium chlorate, and 0.75-4
M sulfuric acid. In about half of the experiments (no.
58-96],we added between ca. 1 and 2.5 M sodium sulfate
in the solution fed to the reactor. In the other runs (no.
1-57), no sodium sulfate was added. Over these ranges we
determined the chlorine dioxide formation rates by
dividing the moles of chlorine dioxide collected in a gas
sample by the time over which the sample was taken and
by the volume of liquid in the reactor. These rates and
the correspondingreactor effluent concentrationsare listed
in Table I. Concentrations of hydrogen peroxide in run
nos. 20 and 21 were too low to analyze accurately by the
permanganate titration procedure. In those two experiments, hydrogen peroxide concentration was estimated
by a material balance, using the hydrogen peroxide feed
rate and concentration,the measured chlorine dioxide rate,
and the stoichiometric relationship of the reaction.
We gathered the batch reactor data at 25-35 "C, 0.05-

1454 Ind. Eng. Chem. Res., Vol. 32, No. 7, 1993


lSZ
1

1012,E= 24 300,x =4.4,y = 0.6,andz = 1.3. Areasonably


I
good fit was obtained as shown by the large t ratios for

0.8

.
U'
U

0.6
0.4
0.2
0

10

20

30

40

50

60

Time, min.

Figure 3. Results of tracer response experiments in flow reactor, for


gas flow rates of 108 (solid symbols) and 0 cms/min (open symbols).
i

-_

0 100

bo

'

10

12

Time, min

Figure4. Resulta of batch reaction experiments at 35 "C, 2 M sodium


chlorate, and 2.6 M sulfuric acid. Symbols refer to concentration of
hydrogen peroxide.

0.75 M hydrogen peroxide, 0.75-2.5 M sodium chlorate,


and 2-3.25 M sulfuric acid. Typical data for several of
these experiments are shown in Figure 4. It shows the
trend of chlorine dioxide vs time which is typical of the
trend in all of the experiments: the data show a transition
period at low time in which rate increases, followed by a
steady-state region. We calculated the chlorine dioxide
rate as the slope of a line through the data in the region
of steady rate. In about half of the batch experiments, we
added 0.5-1 M sodium sulfate to the initial reactant
solution. In the other half, no sodium sulfate was added.
c. Data Correlation: Reaction System without
Added Sodium Sulfate. We correlated reaction data in
terms of a power law model

RCIO,= A exp(-E/RT)[H2S041X[H2021Y[NaC10,12
(5)
using statistical software (MINITAB, 1990). In applying
the program, we fit the data to the logarithmic form of the
equation:
ln(Rclo,) = ln(A) - E/RT + x ln[H2S041+ y ln[H20,1 +
z ln[NaCIOSl (6)
The input consisted of batch and flow reactor concentrations of sulfuric acid, hydrogen peroxide, and sodium
chlorate, reciprocal temperatures, and experimental rates.
In that analysis, the program finds values of the parameters, x , y, z, In A, and EIR, which minimize the difference
between experimental and calculated values of ln(Rcloa)
and performs a statistical analysis of the parameters.
We first fit only those data gathered in experiments in
which sodium sulfate was absent from the feed solution.
The results are shown in Table I1and Figure 5. The kinetic
parameters that best describe these data are A = 4.0 X

each parameter and the adjusted coefficient of determination near unity (Le., R2 = 0.957).
There is some question whether stoichiometric sulfuric
acid concentration is the most appropriate variable to use
in eq 5, although using this variable allows us to compare
our results with those of Hong et al. (1967). They studied
chlorine dioxide production by the chloride-chlorate
reaction and found that the reaction is very highly acid
dependent, with an order of about 13 in [HzS041. By
comparison,we can conclude that the rate of the hydrogen
peroxide-chlorate reaction is much less dependent on
acidity, since we found an order of 4.4 in [H2S041 in the
present study.
d. Data Correlation: System with Added Sulfate.
We found very poor agreement between the rates that
were measured in reaction experimentswhich incorporated
sodium sulfate in the feed solution and the ratescalculated
by eq 5 on the basis of the kinetic parameters in Table I1
(see Figure 6). The deviation between experimental and
calculated rates increased as the concentration of added
sodium sulfate. For most of these experiments, the
measured rates were much lower than the calculated rates.
The poor agreement was expected since sodium sulfate
reacts with protons in sulfuric acid solutions, forming
sodium bisulfate, thereby lowering the acidity of the
solution (Lindstrom and Wirth, 1969). This effect is
partially offset by the fact that sodium ions from the added
sodium sulfate become hydrated in solution, slightly
increasing the acidity of the solution (Hong et al., 1967).
To improve the kinetic model so that it would account
for all of the data of this study, we substituted an acidity
function, h-, in place of [H2S041 ineq 5 to represent acidity
of the solution. Empirical acidity scales are commonly
used to characterize acidity of solutions of high concentration where the pH scale does not apply. The acidity
functions are often tabulated in logarithmic form; for
example, the function h- is most often tabulated as the
function -If-, which equals log(hJ. Tables of acidity
functions vs stoichiometric concentration of sulfuric and
other acids can be found in many references [see, for
example, Boyd (1969) and Rochester (1970)l. These
tabulated functions do not apply to solutions containing
both sulfuric acid and sodium sulfate.
We recently developed a modified version of an Hacidity scale reported earlier by Cox and McTigue (1964).
Our efforts in modifying the scale are described in Indu
et al. (1991b). In essence, this modified scale takes into
account that sodium ions increase acidity and sulfate ions
decrease acidity of sulfuric acid solutions.
Indu et al. (1991b) correlated experimentally measured
acidity functions for sodium sulfate-sulfuric acid solutions
by the equation

-H- lOg(h-) = -2.32

+ 2.64~+ 0.219b - 1.16~'-

0.016b2 + 0.44ab (7)


where a and b describe relative and absolute stoichiometric
amounts of sulfuric acid and sodium sulfate in solution
a=

[H2SO4I/([H2SO41
+ [Na2S041)

(8)

and
b = [H2SO41+ [Na2SO41
(9)
In effect, this equation accounts for the net influence of
sodium ions and sulfate ions contributed by sodium sulfate
to the acidity of sulfuric acid.
For solutions which contain sodium chlorate, Indu et al.
(1991b) showed that the above correlation yields acidities

Ind. Eng. Chem. Res., Vol. 32, No. 7, 1993 1455


Table 11. Results of Multilinear Regression
results for eq 5
parameter value
SD
29.028
1.135
-12236
367
4.4152
0.1126
0.6380
0.0316
1.3069
0.0927

parameter

InA

EIR
It

no. of data points used


e t d model error
coeff of regression

owc
e

9u

%A

.2

-2

-3

-3

-4

-4

I
I

s
U

x Batch

t Plow
Batch, NaZKM added
A Flow, N d O I added

I
-4

-1

Figure 5. Comparison of rates measured in the absence of added


sodium sulfate and rates calculated using eq 5 and the parameters
in Table 11.
-1 1

5'
c
-2

El
Y

33

X Batch

Flow

-3
-2
LOG (EXPERIMENTAL RATE)

.5

4
&

t ratio
28.58
-39.62
48.99
22.60
18.44
153
0.374
0.952

-- I

G
c
&

t ratio
25.59
-33.32
39.21
20.18
14.10
91
0.365
0.957
-1

-1

results for eq 10
parameter value
SD
23.442
0.820
-10505
265
1.8798
0.0384
0.6667
0.0295
1.2642
0.0686

-3

&

Yc

..

Batch, Na2SCU added

4 Flow,

-5

NiZS04 idded

I
-5

.4

-3
-2
LOG (EXPERIMENTAL RATE)

-1

Figure 6. Comparison of rates measured in the presence of added


sulfate and rates calculated using eq 5 and the parameters in Table
11.

which are too low and must be corrected by adding the


quantity 0.12[NaClOs] to the value of -H-determined by
eq 7. In effect, this added quantity corrects for the
influence of sodium ions contributed by the sodium
chlorate. Chlorate ions are believed to have little or no
influence on the acidity (Hong et al., 1967).
Table I shows values of the parameters a and b and the
acidity function -H-for all of the solutions we used in the
present study. To show the influence of sodium ions
contributed by the sodium chlorate on the acidity of the
solutions, values of -H-are tabulated before and after
this sodium ion contribution is corrected for. We used
values of -H-that were corrected in correlating the data
and with the kinetic equation.
Using the MINITAB program, we fit the data gathered
in all of the batch and flow reactor experiments with the
rates predicted by the modified form of eq 5 (with hsubstituted for [HzS041). The results are shown in Table
I1 and Figure 7. The form of the kinetic equation that
best fits the experimental data is

-5

-3
-2
LOG (EXPERIMENTAL RATE)
-4

I
-1

Figure 7. Comparison of experimental rates and rates calculated


using eq 10.

RClO,= 1.5 X 10" exp(-20900/RT)(h~)'~9(H202)0~7


X
(NaC103)'.3 (10)
A reasonably good fit was obtained as shown by the large
t ratios for each parameter and the adjusted coefficient
of determination near unity (Le., R2= 0.95). The statistical
program also performed an analysis of variance, which
showed very high F values, leading t o p values of less than
0.001. This analysis indicates that all four independent
variables have a statistically significant influence on the
rate.

Application and Conclusions


A production trial was conducted at a commercial
chlorine dioxide plant which has the configuration shown
in Figure 1. During a 40-min start-up phase of the trial,
reactor conditions were established and controlled such
that the production data could be compared with the
predictions of eq 10.
The reactor was initially charged with 10000 L of a
solution consisting of 2.94 M sodium chlorate and 1.74 M
sulfuric acid. The solution was heated to boiling and the
pressure controlled to maintain constant temperature at
70 "C. After constant temperature was achieved,reactants
were fed in stoichiometric proportion and maintained at
a constant rate equivalent to a maximum chlorine dioxide
production rate of 10 tons/day (93.55 mol/min). The
reactants were fed as three aqueous solutions: 49 wt %
sodium chlorate; 50 wt % hydrogen peroxide, and 60 wt
% sulfuric acid. Throughout the 40-min period, the reactor
'
solution volume was maintained at the initial value f10 %
by adjusting the steam rate to the recirculation heat
exchanger, which determined the rate of water evaporation.
Also during this period, temperature was maintained at
the initial value f3 OC.
Chlorine dioxide production rate was determined at
regular time intervals by measuring the flow rate and
chlorine dioxide concentration of the aqueous solution
leaving the bottom of the absorber (see Figure 1). These
data are plotted in Figure 8 along with predictions based

1456 Ind. Eng. Chem. Res., Vol. 32, No. 7, 1993

10

8 -

4 -

i-

9u

6 -

2 -

0 -

-2
-10

10

20

30

40

50

Time, min.
Figure 8. Commercial plant production data (TPD, tons per day;
data points represent measured rates, solid line represents predictions
of eqe 10 and 11).

on eq 10 (see solid line in Figure 8). The predictions were


made by numerically integrating the equation
(11)
Vd[i]/dt = mi - yiRcqV
starting with the previously mentioned initial conditions.
There is a fluctuating pattern in the data, perhaps
reflecting adjustments in the process control equipment.
The model based on eqs 10 and 11provides a reasonable
representation of the average trend of the data, although
it underpredicts the production rates at longer times by
about 8%. An increase in the pre-exponential factor of
eq 10 of about 50% would more closely approximate the
data at longer times but would overpredict at shorter times.
In conclusion, the model determined in this laboratory
study using small batch and steady-state flow equipment
reasonably predicts behavior of commercial size equipment. The scale-up factor from laboratory to commercial
scale in this study is between 33 000 and 100 000. On the
basis of these results, we feel that the model should provide
a useful tool for design.

Acknowledgment
We gratefully acknowledge financial support from Eka
Nobel, Inc. We thank the following persons from Eka
Nobel: Mr. John Winters for helpful discussionsregarding
commercial chlorine dioxide processes, Mr. John Sokol
for providing operating data gathered during a commercial
plant start-up, and Mr. StanWeaver for helpful comments
and review of the manuscript.
Nomenclature
A = pre-exponential Arrhenius parameter, M1-x-y- min-
C = sodium carbonate concentration, M
CO= steady-state sodium carbonate concentration, M
E = activation energy, cal mol-
[i] = concentration of reactant i, M
mi = feed rate of reactant i, mol/min
RCQ = chlorine dioxide formation rate, M min-l
R = gas constant, 1.987 cal mol- K-l
t = time, min
T = temperature, K
V = reactor solution volume, L
x y j = reaction orders,respectively, in stoichiometric sulfuric
acid concentration (or in h-), hydrogen peroxide, and
sodium chlorate
yi = ratio of stoichiometric coefficient of species i to that of
chlorine dioxide in eq 3 (1.0 for sodium chlorate; 0.5 for
hydrogen peroxide and sulfuric acid)

Atkinson, E. S. ClOz Generation Cuts Byproduct Sulfuric Acid. Chem.


Eng. 1974,81 (Feb 4), 36.
Boyd, R. H. Acidity Functions. In Solute-Solvent Interactions;
Coetzee, J. F., Ritchie, C. D., Ede.; Dekker: New York, 1969;p 97.
Cox, B. G.; McTigue, P. T. Kinetics of Oxidation of Formate Ions
with Bromine. J. Chem. Soc. 1964,3893.
Emst, W. R.; Shoaei, M.; Fomey, L. Selectivity Behavior of the
Chloride-Chlorate Reaction System in Various Reactor Types.
AIChE J. 1988,34,1927.
Holmstrom, U. K.; Sondgren, L.; Norell, M.; Axegard, P. Process for
Production of Chlorine Dioxide. U.S.Pat. 4,678,664,1987.
Hong, C. C.; Lenzi, F.; Rapson, W. H. The Kinetics and Mechanism
of the Chloride-Chlorate Reaction. Can. J. Chem. Eng. 1967,45,
349.
Hoq, M. F.; Ernst, W. R.; Gelbaum, L. T. NMR Procedure for
Determining Concentrations of Methanol and Formic Acid in
Solutions from a Chlorine Dioxide Generating Plant. Tappi J.
19918,74,217.
Hoq, M. F.; Indu, B.; Ernst, W. R.; Neumann, H. M. Kinetics of the
Reaction of Chlorine with Formic Acid in Aqueous Sulfuric Acid.
J. Phys. Chem. 1991b,95,681.
Hoq, M. F.; Indu, B.; Neumann, H. M.; Ernst, W. R. Influence of
Chloride on the Chlorine-Formic Acid Reaction. J.Phys. Chem.
1991c,95,9023.
Hoq, M. F.; Indu, B.; Emst, W. R.; Gelbaum,L. T. OxidationProducta
of Methanol in Chlorine Dioxide Production. Znd. Eng. Chem.
Res. 1992,31, 1807.
Indu, B.; Hoq, M. F.; Emat, W. R.; Neumann, H. M. Kinetics of the
Reaction of Chlorine with Formaldehyde in Aqueous Sulfuric Acid.
Znd. Eng. Chem. Res. 1991a,30, 1077.
Indu, B.; Hoq, M. F.; Ernst, W. R. Acidity of Sulfuric Acid-Sulfate
Solutions by Kinetic Measurements. AIChE J. 1991b,37,1744.
(In this reference, there is an error in the caption to Figure 4;it
should read Numbers refer to parameter a.)
Indu, B.; Ernst, W. R.; Gelbaum, L. T. Methanol-Formic Acid
Esterification Equilibrium in Sulfuric Acid Solution. J.Org. Chem.
1992,submitted for publication.
Isao ba, M.; Hideo Yamamoto, S. Y.; Syuki Shindo, S.; Morieki
Shibuya, S. Process for Manufacturing Highly Pure Chlorine
Dioxide. U.S.Pat. 4,421,730,1983.
Lindstrom, R. E.; Wirth, H. E. Estimation of the Bisulfate Ion
Dissociation in Solutions of Sulfuric Acid and Sodium Bisulfate.
J.Phys. Chem. 1969,73,218.
Masschelein,W. J. Chlorine Dioxide;Ann Arbor Science: Ann Arbor,
MI, 1979;p 114.
MINITAB. Release 6.2 (copyright 1990); Minitab Inc., 3081 Enterprise Dr., State College, PA 16801.
Norell, M. Process for Production of Chlorine Dioxide. U.S.Pat.
4,770,868,1988.
Rapson, W. H. Recent Developments in the Manufacture of Chlorine
Dioxide. Can. J. Chem. Eng. 1968,36,262.
Rochester, C. H. Acidity Functions; Academic Preee: New York,
1970; p 23.
Roaen, H. J. Method of Simultaneously Producing Chlorine Dioxide
and a Sulfate Salt. US. Pat. 3,933,988,1976.
Smith, J. M. Chemical Engineering Kinetics, 3rd ed.; McGraw-Hilk
New York, 1981;p 275.
Soule, E. C.Manufacture of Chlorine Dioxide. US. Pat. 2,332,181,
1943.
Sprauer, J. W. Production of Chlorine Dioxide. US.Pat. 2,833,624,
1956.
Tenney, J.; Shoaei, M.; Obijeski, T.; Ernst, W. R.; Lindstroem, R.;
Sundblad, B.; Wanngard, J. An Experimental Investigation of the
Chloride-Chlorate Reaction System. Znd. Eng. Chem. Res. 1990,
29,916.
Vogel, A. Textbook of Quantitative Inorganic Analysis, 4th ed.;
revised by Basset, J., Denney, R. C., Jeffrey, G. H., and Mendham,
J.; Longman Group: London, 1978; p 381.

Literature Cited
Aieta, E. M.; Roberts,P. V.; Hernandez, M. Determination of Chlorine
Dioxide, Chlorine, Chlorite, and Chlorate in Water. J.Am. Water
Assoc. 1984,76,64.

Received for review October 5, 1992


Revised manuscript received March 1, 1993
Accepted March 17,1993

You might also like