You are on page 1of 6

Authors Accepted Manuscript

Reply to the comments on the paper titled


Hydrolysis of acetic anhydride: Non-adiabatic
calorimetric determination of kinetics and heat
exchange by Wilson H. Hirota, Rodolfo B.
Rodrigues, Claudia Sayer, Reinaldo Giudici
published in Chemical Engineering Science, 65
(2010) 38493858.
www.elsevier.com/locate/ces

Reinaldo Giudici

PII: S0009-2509(16)00013-0
DOI: http://dx.doi.org/10.1016/j.ces.2016.01.012
Reference: CES12744
To appear in: Chemical Engineering Science
Received date: 20 November 2015
Accepted date: 11 January 2016
Cite this article as: Reinaldo Giudici, Reply to the comments on the paper titled
Hydrolysis of acetic anhydride: Non-adiabatic calorimetric determination of
kinetics and heat exchange by Wilson H. Hirota, Rodolfo B. Rodrigues, Claudia
Sayer, Reinaldo Giudici published in Chemical Engineering Science, 65 (2010)
3 8 4 9 3 8 5 8 . , Chemical Engineering Science,
http://dx.doi.org/10.1016/j.ces.2016.01.012
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
MANUSCRIPT TITLE: Reply to the comments on the paper titled Hydrolysis of acetic
anhydride: Non-adiabatic calorimetric determination of kinetics and heat exchange by Wilson
H. Hirota , Rodolfo B. Rodrigues , Claudia Sayer , Reinaldo Giudici published in Chemical
Engineering Science, 65 (2010) 38493858.

MANUSCRIPT TYPE: letter to the editor

KEY WORDS: Acetic anhydride; Reaction calorimetry; Hydrolysis; Kinetics; Chemical reactors;
Reaction engineering; Mathematical modeling

-------------------------------------------------------------------------------------

Reply to the comments on the paper titled Hydrolysis of acetic anhydride: Non-adiabatic
calorimetric determination of kinetics and heat exchange by Wilson H. Hirota , Rodolfo B.
Rodrigues , Claudia Sayer , Reinaldo Giudici published in Chemical Engineering Science, 65
(2010) 38493858.

Dear Editor,

We appreciate the comments made on our paper Hydrolysis of acetic anhydride: Non-
adiabatic calorimetric determination of kinetics and heat exchange, published in Chemical
Engineering Science, 65 (2010) 38493858. This is an excellent opportunity to revisit the
contribution, to discuss some important points raised and to clarify further some issues. Below
is the response to the comments.

1. Miscibility condition of the initial reaction mixture: We completely agree that the
solubility of acetic anhydride in water at 20C is very low, and that the system is not
completely homogeneous at the initial condition of our experiments. This was indeed
mentioned and treated in our article. However, we completely disagree that the
results obtained would not be valid or reliable. The way we treat the data, this effect
of initial incomplete miscibility does not to affect our results. All the conversion values
(and thus all the reactant concentration at each time) are calculated starting from full
conversion (XA = 1),which was assumed to occur at the very end of the experiment,
and calculated backwards to the initial time (t=0). The true initial condition was
calculated from the results, and was not assumed to be known beforehand. During
the experiment, we could visually observe the reaction medium and see that the
solution was initially turbid (an indication that the system is two-phase). At a slightly
later time, it became completely limpid or transparent (an indication that the system
reached, from this time on, the condition of total miscibility). This happens because,
during our experiment, two factors contribute to reach this miscibility condition: the
temperature gradually increased and, more importantly, acetic acid was being
gradually produced by the reaction. As reported in our paper, this transition to a
completely homogeneous solution typically happened around the time the
temperature reached ~35C, when enough acetic acid was formed (we agree that the
effect of acetic acid is more significant than that of temperature; however, both
factors work in the same direction). Therefore, we were quite aware that, during the
first few minutes of the experiment, we had a two-phase situation. In addition, we
were able to easily identify the point at which the change to a completely miscible
system occurred. Hence, the first few points corresponding to the two-phase period
can be discarded from the kinetic analysis. Moreover, the remaining points, measured
with the system fully miscible, can be safely used in the kinetic treatment. This
procedure was indeed done in our work (as reported in Figure 4 of our article) to
compare the impact on the results of discarding the first points from the kinetic
analysis. Interestingly, we found that discarding these first data points taken during
the initial two-phase system did not change significantly the results for the kinetic
parameters, probably because the quantity of data taken with the fully miscible
system was much greater. This can be clearly seen in Figure 4 of our article. Discarding
the first points would correspond to deleting the rightmost 10 or 15 points in the
Arrhenius plot of our Figure 3; it can be seen that this would not change significantly
the values of the kinetic parameters obtained from the slope and intercept of these
plots. Moreover, when the values of the kinetic parameters were used to simulate the
full experiment from time zero (i.e., including the two-phase period), no large
differences between the model and the experimental data were observed,even in this
initial part (see Figures 5 and 6 of our article). This reinforces our conclusion that this
problem, although existent, has only a marginal effect on the results.However, if one
prefers to take this effect fully into account and discard all the data measured in the
two-phase region, it is sufficient to use only the data taken during the miscible period
and replace the initial condition [t=0, XA(0)=XAo, T(0)=To] by the condition at which the
solution became completely limpid [say, t=t1, XA(t1) and T(t1)]. I would therefore like to
emphasize that our experiments and our proposed data treatment do in fact give valid
results.

2. Purity of the acetic anhydride: We used commercial chemical-grade acetic anhydride


with a typical nominal specification of minimum purity of 97%. After using the same
flask over and over in the lab, the contents can have contact with humid air each time
the flask is opened. This would gradually convert the anhydride into acetic acid, thus
reducing the nominal purity. Because of the reaction with moisture in the air, the main
impurity present in acetic anhydride will be acetic acid. I have used this experiment as
an undergraduate lab class over the last 15 years and have observed that, starting with
a new flask of acetic anhydride at the beginning of the semester, the values of the
initial conversion XAo obtained are around 2-3% (coehent with the nominal purity of
97%). This value increases as the experiment is repeated over and over by different
groups of students, reaching around 6-10% for the last groups at the end of the
semester. The increase in the measured initial conversion XAo is consistent with the
idea of gradual contamination of the acetic anhydride with acetic acid due to the
humidity of the air.
I would like to emphasize here that our data treatment procedure does not require a
very expensive, very pure analytical grade reagent because, no matter what the initial
purity of the acetic anhydride used, the initial conversion XAo is determined from the
experiment.
The catalytic effect of strong acids like hydrochloric acid or sulfuric acid on the
hydrolysis of acetic anhydride is well known and has been reported several times in
the literature. This may explain part of the differences in the kinetic parameters
reported in our Table 8.

3. Initial composition: In our article, we did not assume that at the temperature
maximum attained in the non-adiabatic experiments, the reaction mixture is
completely converted. Nor did we write that. The assumption of complete conversion
at the time of the maximum temperature is not only inaccurate, it is also WRONG. The
calculations reported in our article do not consider this wrong assumption. Instead,
the correct way, effectively used and described in our article, is to consider complete
conversion at the very last measured point of the experiment (at the end of the
experiment, after the reactor was cooled down), which happens far after the instant of
the maximum temperature.
We can easily demonstrate mathematically that the conversion is not complete at the
point of maximum temperature simply by examining the energy balance, as expressed
in eq (6) of our article, reproduced below:

At the maximum temperature point, (dTr/dt)=0, and (Tr-Tamb)0, so necessarily (-rA)0.

4. Reactor wall heat capacity: We agree that the use of a very thin stainless steel reactor
wall could approach better the (implicit) assumption of uniform wall temperature
because the thermal conductivity of stainless steel is about one order of magnitude
higher than that of glass. On the other hand, the total weight of the wall would be
similar to that of a glass reactor, for a typical glassware wall thickness.
In some of our undergraduate class experiments, we have already tested Styrofoam
(expanded polystyrene) cups of the thermal insulating type. In that case, we agree that
the whole wall temperature will not be the same as that of the reaction mixture.
However, the weight of a styrofoam cup is so low that the term (mCp) of the reactor
walls is quite low in comparison to that of the reaction mixture. Overall, there is a
tradeoff between the small heat capacity of the walls and a uniform wall temperature.
The wall dynamic model that takes into account heat losses from the external surface
of the reactor to the atmosphere by natural convection and radiation would be very
complicated to be used together in the analysis of the kinetics. However, such a model
might be very useful to evaluate the effect of the wall material, wall thickness, etc. on
the temperature of the reaction mixture and on the error that could eventually be
propagated to the kinetic analysis.
One should also note that the heat capacity of the impeller, either of the magnetic
type, as used in our article, or of the mechanical type, should also be taken into
account in the total heat capacity of the system. A mechanical impeller can introduce
an additional complication due to temperature non-uniformities (due to heat
conduction through the rod).
The best solution would be to estimate the overall heat capacity of the whole system
(reactor walls, stirring bar, and reactor contents) by using a small electric heater with a
known power dissipation, as is usually employed in calorimetric reactors

5. Computation of the rate constant: The difficulties related to error propagation and
amplification of inaccuracies involved in obtaining derivatives from experimental data
are well known and have been discussed in several references, notably in textbooks on
Chemical Reaction Engineering (e.g., Foglers Elements of Chemical Reaction
Engineering) and numerical techniques (e.g., Himmelblaus Process Analysis by
Statistical Methods). Estimation of derivatives from intrinsically noisy experimental
data is known to be the Achilles heel of the differential method of kinetic analysis. In
order to obtain the derivatives, we typically used the three different techniques
described in Foglers book: finite difference formulae, polynomial adjusting, and the
differentiation by the method of equal areas. All of these have their advantages and
drawbacks. In our article, we reported only the one that is the simplest to apply, the
finite difference formulas. More sophisticated techniques for evaluation of derivatives
from experimental data are welcome to improve the data treatment.

6. Temperature measurement: Improvement in the temperature measurement is


important and valid to insure more precise temperature readings. The temperature
measurements reported in our article were performed with special T-type
thermocouples with an accuracy of 1C (resolution of 0.1C), previously verified
against a calibrated mercury-in-glass thermometer and with standard known
temperature reference baths. We have also used a more accurate RTD (resistance
temperature detector) with similar results. Of course, if the increase in temperature is
not very large in the experiment (e.g., at low acetic anhydride-to-water ratios), the
accuracy of the temperature measurement becomes more and more critical, requiring
more precise measurement devices, such as RTDs and thermistors. Another important
point is to check for the speed of the device response: the thermal inertia of the
temperature measurement device can be easily evaluated by performing step changes
in the temperature of the medium (under the same stirring conditions of the
experiment) to characterize its thermal response and to evaluate whether it is
necessary to make corrections of the temperatures readings during the experimental
run.

7. Activation energy and model validation: The objective of the comparison between the
simulation and the experimental measurements in Figures 5 and 6 of our article is just
to show the coherence of the set of obtained parameters (UA, XAo, ko ,E) adjusted in
the procedure of data treatment. In summary, this comparison only serves to convince
the readers that the whole procedure to obtain the parameters did not
accumulate/propagate experimental errors to the point of damaging the significance
of the parameter values. The experiments reported in our article refer to quite similar
initial conditions (reactor type and reactor geometry were changed in the
experiments). The cross-correlation between the estimates of ko and E from the
Arrhenius equation is well known and can be substantially reduced by replacing ko by
its equivalent k(Tref), were Tref is an average temperature in the range studied.

8. Catalysis by acids: As already mentioned, the catalytic effect of strong acids like
hydrochloric acid or sulfuric acid on the hydrolysis of acetic anhydride is well known
and has been reported in the literature. However, in spite of the rich literature, no full
kinetic model has been proposed that explains fully the kinetics of the hydrolysis of
acetic anhydride under all the varied conditions (different concentrations of the
reactants, different concentrations of external strong-acid catalyst, different ionic
strengths of the reaction medium, solvation and catalytic role of the acetic acid, etc.).
Although the effects of solubility mentioned above did not affect the results of the
particular experiments reported in our article, they are certainly important in the
framework of the kinetic study of the reaction and should be accounted for in future
studies. The combination of different experimental approaches reported in the
literature for measurement the reaction kinetics (conductivity, spectroscopic
techniques, pH, reaction calorimetry, etc.) can contribute to a better understanding of
the hydrolysis of acetic anhydride.

References

W. H. Hirota , R. B. Rodrigues , C. Sayer , R.Giudici Hydrolysis of acetic anhydride: Non-


adiabatic calorimetric determination of kinetics and heat exchange,Chemical Engineering
Science, 65 (2010) 38493858.

H.S. Fogler, Elements of Chemical Reaction Engineering, Prentice Hall, New York (1999)

D.M. Himmelblau, Process Analysis by Statistical Methods, Wiley, New York (1970)

Reinaldo Giudici

Universidade de So Paulo, Escola Politcnica


Department of Chemical Engineering
Av. Prof. Luciano Gualberto, trav. 3, No. 380,
05508-010, So Paulo, SP, Brasil
rgiudici@usp.br

You might also like