You are on page 1of 17

Investigation of the effect of high

temperature exposure on the oxidizing


power of sodium hypochlorite bleach

AUTHOR

Kai Roger Smith Hove

SUPERVISOR

Anniken T. vers

SCHOOL

Arendal Videregende Skole

COMPLETED

05.11.2011

WORD COUNT

3978 words

Abstract
Liquid bleach, in the form of sodium hypochlorite (NaClO), is an unstable compound,
and is subject to a number of decomposition pathways depending on pH, UV
radiation, transition metal contamination and temperature. The scope of this
experiment includes an investigation of the effect of the latter: How does exposure to
high temperatures affect the oxidizing power of sodium hypochlorite bleach (on
iodide)? A hypothesis predicting reduced oxidizing power at higher temperatures was
made on the grounds of a proposed thermal decomposition pathway, which predicts
the formation of chlorate and chloride salts. A high concentration of the chlorate ion,
which displays a weak standard reduction potential, would reduce the oxidizing power
of the solution.
In the experiment, four undiluted samples of commercial bleach solution were heated
within a temperature range of 30-40 C, 50-60 C, 70-80 C and 90-100 C
respectively, over a period of 2 hours, with a reflux condenser. After allowing cooling
overnight, a series of iodometric titrations were performed to determine the oxidizing
power of the solutions. This involved the oxidation of excess iodide to iodine by
hypochlorite in acidic media. The liberated iodine was then reduced back to iodide by
titration with sodium thiosulfate with a starch indicator. The oxidizing power of the
bleach was subsequently extrapolated with respect to how much iodide that was
oxidized under each temperature condition.
Unfortunately, no strongly justified conclusions could be drawn due to the significant
spread in the data, which indicated the presence of a systematic error. However,
qualitative data in the form of crystallization of a hygroscopic substance was observed
in the two samples heated above 70 C, something which partly supports the initially
proposed chlorate-forming decomposition pathway.
Words: 294

Table of contents

Section

Title

Page

Abstract
1

Introduction and research question

Background theory

Hypothesis

Method

4.1

Heating the bleach

4.2

Iodometry

Results

Analysis

10

Conclusion

11

Evaluation

14

8.1

Error sources and improvements in experimental design

12

8.2

Literature evaluation

13

Unanswered questions and further research


Bibliography

13

1 Introduction and research question


Sodium hypochlorite (NaClO), was first synthesized in the late 1800s through the efforts
of French chemists Claude Louis Berthollet and Antoine Labarraque. It came into
widespread use in the 1930s, and to this day it remains a high-demand product. Its
desirable oxidizing and antibacterial properties lend themselves to a wide range of
applications, from many industrial and professional uses such as in endodontics
(Clarkson, 1998), to general household cleaning and bleaching of clothes.
However, the use of sodium hypochlorite comes with an inherent drawback; it is an
unstable compound, and is subject to degredation through various decomposition
pathways catalyzed by factors such as UV radiation from sunlight, transition metal
contamination and temperature. (Harms et. al., 2010) The scope of this investigation
includes the effect of the latter.
For both producers and end users of bleach alike, it is important knowing to what extent
high temperature exposure affects the oxidizing power of the bleach solution, and in light
of this knowledge, handle the solution correctly with respect to its production, storage
and use in order to maximize its efficiency.
The oxidizing power of the bleach solution can be inferred from to the extent to which it
can successfully oxidize other compounds. In this investigation, iodide has been used as
a reducing agent.

This leads to the research question of this essay: How does exposure to high
temperatures affect the oxidizing power of sodium hypochlorite bleach on iodide?
During preliminary investigation of the subject, it would appear that previous research on
the effect of temperature on sodium hypochlorite solutions has yielded contradictory
results, something that was also noted by Frais et. al. (2001). While some researchers
have reported little to no effect of high temperatures on the oxidizing efficacy of sodium
hypochlorite solutions (Gambarini, G., De Luca, M. & Gerosa, R., 1998), Frais et. al.
reported a significant loss in oxidizing efficacy. This unresolved area of research was
something that peaked my personal interest, and why I deemed it worthy of further
investigation.

2 Background theory
Commercial bleach solution is produced by bubbling chlorine gas through an aqueous
solution of excess sodium hydroxide. The resulting solution of bleach consists of sodium
hypochlorite, sodium chloride, water, as well as excess sodium hydroxide. (Harms et. al.,
2010)
!

Cl2 (g) + 2 NaOH (aq) NaClO (aq) + NaCl (aq) + H2O (l)

In the precense of water, the sodium hypochlorite will then dissociate into ions,
!

NaClO (aq) Na+ (aq) + ClO (aq)

whereas the hypochlorite ion ClO is responsible for the oxidizing action of the bleach,
while the sodium is simply a spectator ion.
As previously mentioned, the problem with liquid sodium hypochlorite in solution is that it
has a tendency to decompose through reactions with itself. Sodium hypochlorite is subject
to several decomposition pathways, of which the primary is a reaction where three moles
of hypochlorite ions react to form one mole of chlorate and two of chloride. (Zoller, 2009)
This is a disproportionation reaction, as Cl is simultaneously oxidized and reduced.
!

3 ClO (aq) 2 Cl (aq) + ClO3 (aq)

Another, much less prevalent way of degradation is the oxygen-producing decomposition


pathway, which results in the formation of sodium chloride and oxygen gas. (Lister, 1956)
!

2 OCl (aq) 2 Cl (aq) + O2 (g)

This occurs to a negligable extent and has not been proven to be temperature-dependent,
but the presence of certain transition metal ions, namely copper, nickel, iron, cobalt and
manganese, have been shown to greatly accelerate this reaction through catalytic activity.
(Lister, 1956) Normally, under proper bleach production conditions, the amount of these
ions would be so insignificant as to not catalyze any significant decomposition. For that
reason, the discussion in this essay will take place under the assumption that chlorate
formation is the main decomposition pathway.
The problem that arises is the conflicting presences of the hypochlorite ion, ClO, versus
the chlorate ion, ClO3. When we are discussing in terms of the oxidizing efficacy of these
ions in bleach solution, which is the more desirable oxidant? There appears to be
conflicting theory surrounding this problem:
ClO3 is a highly oxidized anion, displaying chlorine in the +5 oxidation state, versus ClO,
which is +1. As as ClO3 displays a higher oxidation state, it follows that it should also be
expected to show stronger oxidizing properties. However, this does not seem to be the
case.

The following table, adapted from Cotton et. al. (1988), shows the standard reduction
potential in volts for the successive oxyacids of chlorine through their reduction in acidic
media. Each species has its own standard reduction potential, which is a measure of the
species affinity for electrons - with greater potential, it shows a greater tendency to be
reduced.
Table 1; Standard electrode potentions in volts for the reduction
of successive oxyacids of chlorine in acidic media
Reduction half-reaction

E (V)

H+ + HClO + e Cl2 + H2O

+1.63

3H+ + HClO2 + 3e Cl2 + 2H2O

+1.64

6H+ + ClO3 + 5e Cl2 + 3H2O

+1.47

8H+ + ClO4 + 7e Cl2 + 4H2O

+1.42

From this data, we can see that the hypochlorite ion displays the strongest tendency to be
reduced, while this trend decreases with the successive oxychlorines (with the exception of the
chlorite ion, ClO2, which is a marginally stronger oxidizer than its predecessor). This trend of
weaker oxidizing power in relation to higher oxygen content seems paradoxical, considering it
conflicts with the usual trend of higher oxidizing potential in relation to higher oxidation states.
One possible explanation for this discrepancy relates to the higher ions increased kinetic
stability. Chlorates relative ineffectiveness as an oxidizer compared to hypochlorite could stem
from the increased number of electrons involved in the formation of -bonds in the transition
from ClO to ClO3. (Tyagi, 2009) This, combined with the shielding of the central chlorine atom
by the increased number of oxygens, would add to the ions stability, making it more
unreactive. The Lewis structures of the ions are illustrated below.

Figure 1
Lewis structures of hypochlorite and chlorate

Hypochlorite

Chlorate

-bonds

3 Hypothesis
My hypothesis is that as higher temperatures will result in an increase in kinetic energy, this
should increase the rate of the reaction where hypochlorite decomposes to chlorate. Chlorate being the less effective oxidizer, as demonstrated by its lower reduction potential and more
stable structure - should oxidize less iodide. In other words, I expect heated bleach to be less
effective.

4 Method

The first step of this experiment involves the heating of the bleach. As perfectly accurate
temperature control is not possible with the available equipment, four ranges of temperature
will be employed instead: 30-40 C, 50-60 C, 70-80 C and 90-100 C. Furthermore, high
temperatures up to 100 C was chosen because sodium hypochlorite decomposition is usually
a very slow process that happens over several months time under room temperature
conditions. Thus, higher temperatures should be able to accelerate decomposition enough to
produce quantitiatively significant results within the time constraints of the investigation. A
heating time of 2 hours was employed. The method of heating was partially adapted from the
method by Frais. et. al. (2001). As opposed to their method, which consisted of heating the
bleach uncovered, I will be utilizing a reflux condenser to retain the concentration of the
solutions, which is an important control variable in this experiment.
As previously mentioned, decomposition of hypochlorite largely results in the formation of
chlorate ions. In light of this, a very accurate method of measuring the extent of bleach
decomposition could be through the detection of chlorate ions by ion chromatography or NMR
spectra. Consequently, the oxidizing power could be extrapolated from the chlorate ion
concentration, which should be inversely related. However, while this method would provide
us with very accurate results, such sophisticated equipment is not readily available. Instead, a
simpler, though admittedly less accurate method can be used to indicate the oxidizing power
of the solutions iodometry.
In acidic solution, hypochlorite ions in the bleach will oxidize iodide ions to form iodine,
chloride ions and water.
!

ClO (aq) + 2 I (aq) + 2 H+ (aq) I2 (aq) + Cl (aq) + H2O (l)

The liberated iodine forms an equilibrium with itself, giving rise to triiodide ions, which turn the
solution distinctly yellow:
!

I2 (aq) + I (aq) I3 (aq)

The solution can then be titrated with sodium thiosulfate. The thiosulfate ion reduces the
iodine back to iodide, gradually draining the solution of its color in the process.
!

2 S2O32 (aq) + I2 (aq) 2 I (aq) + S4O62 (aq)

With the solution gradually taking on only a faint yellow color, the endpoint can be difficult to
determine. For that reason, starch solution is added. The iodine forms a reversible, deeply
dark-blue colored complex with the starch that is useful for accurate endpoint detection. When
the dark blue color disappears, this indicates that all the iodine has been reduced back to
iodide ions. From the molarity of the titrant added and the reaction stoichiometry, it is possible
to deduce how many moles of iodide were oxidized by the hypochlorite. This would provide an
indication of the oxidizing power of the solution.
In this experiment I expect thermally decomposed bleach would contain higher chlorate
content. Chlorate, being the weaker oxidant, should oxidize less iodide ions.

Table 2; Variables

Variable type

Variable

Method of measurement /
control

Oxidizing power

Indicated by amount of iodine


liberated through oxidation,
determined through iodometric
titration

Temperature

Record using a temperature node


attached to a digital device and
lowered into solution.
Temperature controlled through
adjusting the electric heating
mantles.

Type of bleach

Klorin, produced by Lilleborg A/S.

Volume of bleach

50.0 mL will be heated.

Concentration of bleach

Utilize reflux condenser in order


to recondense vapors when
heating. Extracted samples will be
diluted by the same amount
before titration.

Heating time

All four samples will be heated for


2 hours.

Na2S2O3 and KI
solutions

Na2S2O3, the titrant, can


decompose over time, and KI can
be gradually oxidized by air,
volatilizing the iodine. Therefore,
these solutions will be prepared
from solids before use.

Addition of starch

Must happen close to the end


point, otherwise the iodine can
form an irreversible complex with
the starch.

Dependent

Inependent

Control

4.1 Heating the bleach

For the process of heating, four round-bottom boiling flasks were each filled with 50.0 mL of
undiluted bleach, measured out in a measuring cylinder. A reflux condenser was utilized in
order to prevent evaporation. The flask and tube was put on top of an electric heating mantle
and secured with a clamp. The setup was placed in the fume hood, as to make away with
what small amount of vapors that may escape. A temperature node attached to a digital
PASPORT Xplorer GLX device was lowered into the flask, and porcelain boiling chips were
added to prevent sudden, uncontrolled boiling. The heat was then turned on for all flasks, and
the time was recorded from the point when all the solutions were in their designated
temperature range, which took roughly 8 minutes. The solutions were heated for 2 hours
before the heat was turned off. Interestingly, hints of a crystalline precipitate had formed in the
solutions heated above 70 C. The samples were left in the fume hood overnight to restabilize
to ambient temperature and for vapors to recondense.

Figure 2
Apparatus setup

Electric heating mantle

4.2 Iodometry

The following day, the bleach solutions were dismounted and remeasured and were found to
have lost <1 mL of water, something which should not impart a significant influence on the
results.
4.96 g of sodium thiosulfate pentahydrate was dissolved in 1.00 L of distilled water to make a
solution of 0.0200 M Na2S2O3. 1.66 g of potassium iodide was dissolved in 100 mL of water to
make a solution of 0.100 M KI. A 2.00 g/L starch solution was prepared under heating until
complete dissolution.
10.0 mL of each of the bleach solutions was pipetted into to a volumetric flask, diluted to 250
mL with distilled water and homogenized. Followingly, a 5.00 mL aliquot of the dilute solution
was pipetted into an Erlenmeyer flask and diluted with another 20.0 mL of water, measured
out in a 25.0 mL measuring cylinder, to increase the volume. 2.00 mL of a 6.00 M stock
solution of HCl was added to the titration to flask to provide the acidic environment needed for
the oxidation reaction to proceed. After this, 2.00 mL of the KI solution was added, initiating
the oxidation reaction between ClO and l. Rapidly after, the solution was titrated against
Na2S2O3 from the burette until the solution was only faintly yellow, after which 2.00 mL of the
starch solution was added, turning the solution strongly blue. Titration was continued with
consistent swirling of the flask until the solution just turned colorless, indicating complete
reduction of the iodide.

5 Results

Quantitative raw data


Table 3; Volume of Na2S2O3 required to reach endpoint in iodometric titration
Vol. Na2S2O3 in mL 0.05
Trial 1
Trial 2
Trial 3

30-40 C
12.40
11.00
11.05

50-60 C
8.90
9.30
8.95

70-80 C
8.90
10.05
11.35

90-100 C
5.85
5.40
8.55

Control (RT)
7.40
5.50
6.10

Qualitative raw data


After one hour of heating, a white, crystalline substance had accumulated at the bottom
of the flasks for temperature ranges 70-80 C and 90-100 C. After the powder was
extracted from another sample and left to dry in air, it started visibly turning into a
greyish slurry.
As for the iodometry; after adding KI to the bleach solution, it turned yellow. Following
titration with Na2S2O3, the color faded until it was only faintly yellow - at which point
starch was added, and the solution turned black-blue. Titration was continued until it
turned colorless. Several minutes after the titration, the solution had again taken on a
very faint blue tint.

Processed data
The following table shows the calculated number of mols for each reaction. The
number of mols is the amount of sodium thiosulfate required to completely reduce the
iodine that resulted from the hypochlorite oxidation of iodide. To reduce random error,
the mole calculations were based on an average of three parallell trials for each
temperature range.

Table 4; Processed data


Vol. Na2S2O3 in mL 0.05
Trial 1
Trial 2
Trial 3
Spread
Average
mols Na2S2O3 / I-

30-40 C
12.40
11.00
11.05

50-60 C
8.90
9.30
8.95

70-80 C
8.90
10.05
11.35

90-100 C
5.85
5.40
8.55

Control
7.40
5.50
6.10

1.40
11.50

0.40
9.10

2.45
10.10

3.15
6.60

1.90
6.30

2.300 10

1.840 10

2.202 10

1.320 10

1.260 10

The number of mols of oxidized iodide was calculated followingly, based on the known
volume and concentration of the sodium thiosulfate solution and the stoichiometry of the
reaction equations.

Example calculation
n(Na2S2O3) = cV = 0.0200 mol dm x 0.0115 dm
n(Na2S2O3) = n(I) = 2.300 10 mol

The results are presented in the following diagram:


Figure 3
Mols Ioxidized

3.000 10
2.700 10
2.400 10
2.100 10
1.800 10
1.500 10
1.200 10
9.000 10
6.000 10
3.000 10
Control (RT*)

30-40

50-60

* room temperature

70-80

90-100

Temperature
range in C

6 Analysis
We would have expected the solutions that were exposed to higher temperatures to have
undergone a faster rate of decomposition to chlorates and chlorides, thus reducing the
amount of hypochlorite available for oxidation, resulting in less oxidized iodide, and
ultimately requiring less volumes of titrant to completely reduce the oxidation products.
Disappointingly, it appears that the experiment was not a success. The data cannot be
interpreted with respect to the hypothesis, as the results do not show a consistent trend,
and a significant spread can be observed even between parallel trials. This suggests the
precense of a systematic error.
There are two major discrepansies that discredit the recorded data: i) there is a large
spread across most parallell trials, the largest being a difference of a whole 3.15 mL of
titrant in the 90-100 C sample, and ii) the control sample, stored at ambient conditions,
displays the lowest oxidizing power of all the solutions, which is the opposite of what is
expected.
Some parallel trials show a gradual positive increase in the titrant added (90-100 C),
some a negative one (30-40 C), and one remained relatively consistent across parallel
trials (50-60 C). They are certainly not uniform as they should be - preferably, titrations
should agree to the nearest drop. It seems the direction of the error goes both ways,
suggesting that the cause of the error present is of a fluctuating nature. For that reason it
is unlikely that the error is related to the standardized thiosulfate solution - if there was a
problem relating to its quality the error should have been in a uniform direction. Other
external factors must have played a role, which are discussed under section 8.

10

7 Conclusion
In light of the inconsistensies of the quantitative data, giving a definite answer is difficult,
as they do not display any consistent trends. In all likelihood, this was caused by a yet
unidentified systematic error. As such, we cannot make a justified assertion as to
whether exposure to high temperatures affects the oxidizing power of sodium
hypochlorite bleach, and the hypothesis cannot be accepted. However, some qualitative
data supports the hypothesis. While the qualitative data is not sufficient in itself, it can
be taken into consideration and contexualized in relation to other research.
In the bleach samples of 70 C and above, visible salt crystallization had occured
most likely a mixture of sodium chlorate and sodium chloride, as per the proposed main
decomposition pathway. This suggests the solution was sufficiently decomposed as to
be saturated with ions. A significant precense of NaClO3 could be supported by the
observation that the salt was highly hygroscopic, as it turned into a slurry over time. This
is a property that characterizes NaClO3. (Eagleson, 1988) This suggests a considerable
precense of chlorate ions, but it remains unknown whether this correlates with the
solutions oxidizing power.
The data is partially supportive of the results found by Frais et. al. (2001). With a
storage temperature of 37 C during the course of 6 months, their bleach sample was
found to have experienced a 62% reduction in chlorine species available for oxidizing
action. Solid salt formation had also occured in their sample, which is also attributed to
be a mixture of sodium chlorate and sodium chloride, but this was not experimentally
confirmed.
With more parallell trials and the implementation of the improvements discussed under
section 8, it should be possible to make valid observations in this experiment.

11

8 Evaluation
8.1 Error sources and improvements
in experimental design
Seeing as errors appeared to occur in a random direction, this makes it is difficult to
pinpoint one defining source of error. A list of errors and their potential
improvements will be discussed.
A major weakness of the experiment was that it was only performed once.
Therefore, regardless of the direction of the data, there may not be strong enough
support to claim that the results show a consistent pattern. A repetition of the
procedure at least three times could provide grounds for stronger evidence. Also,
doing more than three parallell runs for each titration, preferably around ten, would
greatly help reduce random error.
As for the iodometry, a possible source of systematic error includes oxidation of the
iodide ions by O2 in the air. Considering the relatively short time taken to complete
a titration it may not exert a big difference (weak reoxidation was only visibly
apparent 10-15 minutes after reduction with thiosulfate), but it is a probable error
source that should be limited nonetheless. One way of doing this could be by filling
the titration flask with a less reactive gas, like N2 or CO2, during the titration,
something which would displace the oxygen and limit the influence of O2 oxidation.
Another area of improvement is the temperature control under heating. The flasks
with solutions were heated directly in electric heating mantles, which proved to not
result in very precise temperature control. Within very small variations on the heat
setting, temperature was prone to fluctuate within a spread of 10 C. For that
reason, temperature ranges had to be chosen as opposed to specific temperatures.
This limits the precision of the data. With more precise temperature control, which
could be accomplished through slower, more deliberate heating over a longer
period of time, this would reduce the influence of random error and allow for more
justified statements regarding the effect of a particular temperature on the solution.
An extensive treatment of quantitative uncertainties in measurements was omitted.
This could be improved by taking into account all uncertainties and calculating the
percentage uncertainty.
An inherent limitation of the procedure is that we cannot ascribe the oxidizing action
to a specific ion - we can only draw conclusions based on the theories of the ions
structures and reduction potentials. The oxyanions of chlorine display similiar
chemical properties and are prone to rapid interconversion, making it difficult to
isolate or distinguish between them by chemical tests. This limitation could be
overcome by using more sophisticated techniques; the relative presences of certain
anions is frequently measured using ion chromatography. With this method, the
samples could be double checked to see if the hypochlorite anion concentration
does in fact correlate with the solutions oxidizing power. Another method could
involve the use of nuclear magnetic resonance spectra, calibrated for detecting the
presence of the different ions.

12

8.2 Literature evaluation


The works cited in this investigation are taken from, through my best judgment,
credible authors under established publishers, academic journals and reviews from
industrial research institutions. Therefore, I assume with fair certainty that the
essential literature consulted is accurate.

9 Unanswered questions and further research


Some unanswered questions remaining are:

What is the ratio of chloride to chlorate ions in the solution before and after heating?
What is the oxidation-reduction potential of the solution before and after heating?
(This could be determined using an ORP meter.)
Does the duration and intensity of heating affect how the bleach decomposes? Short
and intense exposure to heat or long exposure at a mild temperature?
Can bleach solutions be effectively preserved at cold temperatures over longer time
periods?

The experiment leaves multiple opportunities for further research. Since we are
discussing temperature, one could also approach an investigation from a kinetic
viewpoint:

What is the rate and reaction order of the decomposition process?


Does the reaction satisfy the rule of thumb for a doubling in rate for every 10 C
increase in temperature? (Brown, C. & Ford, M., 2009)

Other possible areas of research include the reaction mechanism for transition metal
catalyzed decomposition, and the effect of UV radiation (in the form of sunlight exposure,
for instance). This combined knowledge should provide a good foundation for
understanding how sodium hypochlorite bleach should be produced and stored in order
to maximize its oxidizing efficacy.

13

Bibliography

Brown, C. & Ford, M. (2009) Higher Level Chemistry for the IB Diploma. Essex:
Pearson Education Ltd. 263

Clarkson, R. M. & Moule, A. J. (1998) Sodium hypochlorite and its use as an


endodontic irrigant. Australian Dental Journal, 43(3), 250-256. DOI: 10.1111/j.
1834-7819.1998.tb00173.x

Cotton, F. A., et. al. (1988) Advanced Inorganic Chemistry, 5th edition. New
York: John Wiley & Sons. 564

Eagleson, M. (1994) Concise Encyclopedia Chemistry. Berlin: Walter de Gruyter.


996

Frais, S., Ng, Y-L. & Gulabivala, K. (2001) "Some Factors Affecting the
Concentration of Available Chlorine in Commercial Sources of Sodium
Hypochlorite". International Endodontic Journal, 34(3), 206-215. DOI:10.1046/
j.1365-2591.2001.00371.x

Gambarini, G., De Luca, M. & Gerosa, R. (1998) Chemical stability of heated


sodium hypochlorite endodontic irrigants. J Endod, 24(6), 432-434. DOI:
10.1016/S0099-2399(98)80027-7

Harms, L. L. et al. (2010) White's Handbook of Chlorination and Alternative


Disinfectants, 5th Edition. New York: John Wiley & Sons. 468-473

Lister, M. W. (1956) Decomposition of Sodium Hypochlorite: The Catalyzed


Reaction NRC Research Press. Retrieved 21.05.2011 from
http://www.nrcresearchpress.com/doi/pdf/10.1139/v56-069.

Tyagi, V. P. (2009) Essential Chemistry. Delhi: Ratna Sagar P. Ltd. 7.87

Wall, D. (2002) "Chemical Behavior of Hypochlorite in High Ionic Strength


Solutions." Sandia National Laboratories. Retrieved 21.05.2011 from
http://www.nwmp.sandia.gov/onlinedocuments/wipp/tp/tp0201.pdf

Zoller, U. (2009) Handbook of Detergents, Part F: Production. Surfactant


Science Series, vol. 142. Boca Raton: CRC Press. 435

You might also like