Sei sulla pagina 1di 10

ARTICLE IN PRESS

Chemical Engineering Science 65 (2010) 38493858

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Hydrolysis of acetic anhydride: Non-adiabatic calorimetric determination of


kinetics and heat exchange
Wilson H. Hirota a, Rodolfo B. Rodrigues a, Claudia Sayer b,1, Reinaldo Giudici a,
a
b

cnica, Departamento de Engenharia Qumica, Av. Prof. Luciano Gualberto, trav. 3, n. 380, CEP 05508-900 Sa~ o Paulo, Brazil
Universidade de Sa~ o Paulo, Escola Polite
polis, Brazil
Universidade Federal de Santa Catarina, Departamento de Engenharia Qumica e de Alimentos, Floriano

a r t i c l e in fo

abstract

Article history:
Received 14 August 2009
Received in revised form
14 March 2010
Accepted 19 March 2010
Available online 24 March 2010

A simple calorimetric method to estimate both kinetics and heat transfer coefcients using
temperature-versus-time data under non-adiabatic conditions is described for the reaction of
hydrolysis of acetic anhydride. The methodology is applied to three simple laboratory-scale reactors
in a very simple experimental setup that can be easily implemented. The quality of the experimental
results was veried by comparing them with literature values and with predicted values obtained by
energy balance. The comparison shows that the experimental kinetic parameters do not agree exactly
with those reported in the literature, but provide a good agreement between predicted and
experimental data of temperature and conversion. The differences observed between the activation
energy obtained and the values reported in the literature can be ascribed to differences in anhydride-towater ratios (anhydride concentrations).
& 2010 Elsevier Ltd. All rights reserved.

Keywords:
Acetic anhydride
Reaction calorimetry
Hydrolysis
Kinetics
Chemical reactors
Reaction engineering
Mathematical modeling

1. Introduction

anhydride can be represented as


O

During the early stages of a chemical process development,


tasks such as delineating safety conditions and reaction optimization and control can be carried out only if a reaction model and
the corresponding reaction parameters are known. If the kinetics
is not fully understood, several problems may arise, including
runaway reactions that can lead to the release of environmentally
dangerous substances, serious physical equipment damage, or
even the possibility of an explosion. Unfortunately, for the
majority of reactions, there is hardly any knowledge about the
kinetic, thermodynamic, and physical parameters.
Since several chemical reactions are moderatelyhighly
exothermic, it is possible to quantify continuously the amount
of heat released based only on temperature measurements and
energy balance equations that, in turn, can be used to infer useful
information about the progress of the reaction such as thermochemical and kinetics parameters, reaction calorimetry becoming
an essential tool for data-oriented development of chemical
engineering processes.
Hydrolysis of acetic anhydride is a moderatelyhighly exothermic,
fast reaction, which is ideal for verifying the dynamic response of
a calorimetric reactor. The overall hydrolysis reaction of acetic

O
H3C

O
O

CH3
CH3

CH3

H2O

CH3
C

O
H

elimination
CH3

CH3

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.03.028

H2O

OH

The overall reaction mechanism proceeds via three irreversible


steps: addition, elimination, and proton transfer to solvent
(Asprey et al., 1996):
addition

O
H

CH3

 Corresponding author. Fax: +55 11 3813 2380.

E-mail address: rgiudici@usp.br (R. Giudici).


Fax: +55 48 3331 9687.

H3C

H3C

CH3
C

O
H

ARTICLE IN PRESS
3850

W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

proton transfer

where the addition reaction is the rate-controlling process, and


the resulting kinetics are rst-order in each of the reactants
(Asprey et al., 1996).
The kinetics of the hydrolysis of acetic anhydride have been
studied in literature using different techniques for measuring the
extent of the reaction and the reaction rate: titration (Orton and
Jones, 1912; Cleland and Wilhelm, 1956), colorimetry (Oakenfull,
1974), conductivity (Rivett and Sidgwick, 1910; Asprey et al.,
1996; Kralj, 2007), spectroscopic techniques (Bell et al., 1998;
Zogg et al., 2004; Haji and Erkey, 2005), and different calorimetric
techniques (Gold, 1948; Smith, 1955; Janssen et al., 1957; Glasser
and Williams, 1971; Regenass, 1985; Shatynski and Hanesian,
1993; Ampelli et al., 2003, 2005). More recently, the combination
of calorimetric and spectroscopic techniques was also proposed
(Zogg et al., 2003, 2004; Puxty et al., 2005).
Conn et al. (1942) employed calorimetry to measure the
enthalpy of hydrolytic reactions at 303 K for a number of straightchain and cyclic acid anhydrides (acetic, propionic, isobutyric, and
trimethylacetic anhydrides). Gold (1948) analyzed the effect of
solvent type (water and acetonewater mixtures) and temperature on the kinetics of the hydrolysis of acetic anhydride; rstorder rate constants were determined by conductimetry at 5, 15,
and 25 1C, and the data analyzed in terms of the parameters of the
Arrhenius equation; the reaction belongs to the class of
nucleophilic bimolecular (SN2) solvolytic substitution reactions
and, therefore, the hydrolysis of acetic anhydride obeys second
order kinetics, being rst order with respect to both acetic
anhydride and water. In order to elucidate the role of water
structure in the kinetics of hydrolysis, Oakenfull (1974) made a
comparative study of the kinetics of hydrolysis of acetic
anhydride and the reaction of 4-nitrophenyl acetate with
imidazole in mixtures of water with ethanol, t-butyl alcohol,
dimethyl sulphoxide, and dioxin; he observed that both rate
constants were always reduced by the addition of organic solvent.
Janssen et al. (1957) used calorimetric techniques to study the
hydrolysis of acetic anhydride in concentrated acetic acid without
catalyst.
Dyne et al. (1967) and King and Glasser (1965) measured the
kinetics of hydrolysis of acetic anhydride in diluted solutions
using an adiabatic calorimetric reactor in which the reactor walls
were electrically heated to be at the same temperature of the uid
inside the reactor. Later, Glasser and Williams (1971) determined
the kinetic parameters of the hydrolysis of acetic anhydride in
dilute aqueous solution, from experimental temperaturetime
curves by a regression analysis on the differential equations
describing the reaction system. A small vessel reactor, working as
an isoperibolic calorimeter, was placed in a constant-temperature
bath, and the heat transfer coefcient between reactor and the
bath was determined from temperaturetime curve obtained
using a known amount of a hot or cold suitable inert liquid added
in the reactor using a syringe lled with a known quantity of one
reactant. In order to eliminate the mixing effects when the
reactants are rst brought together, the zero of the time scale of
differential equations that describe the reactions system was
chosen at an arbitrary point after the initial disturbances died
away. Shatynski and Hanesian (1993) also determined the
kinetics of the hydrolysis reactions using temperature data

obtained under adiabatic conditions in a commercial reaction


calorimeter in order to avoid some of the complexities inherent in
the method proposed by Glasser and Williams (1971).
Asprey et al. (1996) used the temperature scanning method to
estimate the kinetic parameters of a simple reaction of acetic
anhydride hydrolysis carried out in a plug-ow reactor (PFR),
continuous stirred tank reactor (CSTR), and batch reactor. The
conversion of acetic anhydride was calculated from conductivity
measurements using available computer interface boards. Some
noise in the conductivity signal was unavoidably present due to
the electrodes used. The operation of a temperature scanning
reactor involves the ramping of the feed temperature while
collecting composition and temperature data at the output of the
reactor. In order to make such data interpretable, the temperature
ramping rate must be such that the time between successive
analyses is much shorter than the time required for a kinetically
signicant increment of the temperature to be induced in the feed
to the reactor (Wojciechowski, 1997). As the reactions were
carried out with an excess of water, pseudo-rst-order reaction
conditions were assumed. Comparison of Arrhenius plot determined by temperature scanning with those found in the literature
showed that the reaction constant obtained by this method is
smaller than those obtained by the methods previously used in
studies of acetic anhydride hydrolysis.
Bell et al. (1998) monitored the hydrolysis of acetic anhydride
in a hydrothermal/supercritical water reactor with Raman
spectroscopy to demonstrate the use of in situ spectroscopy
combined with mathematical interpretation to estimate kinetics
and thermodynamics parameters of an organic reaction carried
out in a challenging medium. From the intensity relationships and
the temperature information, the activation energy of the
hydrolysis reaction was estimated and compared to literature
values. Besides, the authors also estimated the enthalpic and
entropic cost of this reaction for a change in solvation environment of acetic acid at high temperatures. The authors noted that
the value of activation energy obtained by this method is
consistent with the literature value of acid-catalyzed hydrolysis,
but signicantly lower than those values reported for noncatalyzed process. The experiments were carried out in a stainless
steel high-pressure reactor vessel with internal volume of 10.4 mL
using 2 mL of distilled water and 2 mL of acetic anhydride.
Zogg et al. (2004) proposed a new approach that combines the
joint evaluation of calorimetric and online infrared data to
identify kinetic and thermodynamic parameters of a chemical
reaction. Furthermore, a new weighting principle was developed
that performs an automatic scaling of the infrared and calorimetric data in order to equalize their inuence on the estimated
reaction parameters. According to the authors, the evaluation of
the calorimetric and infrared data is conventionally carried out
separately. This will generally lead to different estimated reaction
parameters caused by measurement errors, by unmodeled
processes with different inuences on the two analytical signals,
or unequal information content of the two signals. The feasibility
of the new evaluation algorithm was demonstrated based on the
hydrolysis of acetic anhydride. The authors noted that the
reaction of acetic anhydride with water shows signicant heat
of mixing, this effect is visible only in the calorimetric signals and
does not affect the infrared spectrum, and, therefore, the
separated evaluation of the infrared and calorimetric data gave
similar results only when the dosing phase was excluded from the
calorimetric data. As the reaction was carried out in acid medium,
the reactions kinetics was assumed to be pseudo-rst-order. Zogg
et al. (2003) and Visentin et al. (2004) developed a small-scale
reaction calorimeter tted with integrated infrared-attenuated
total reection (IR-ATR), which combines the principles of power
compensation and heat balance. The power compensation was

ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

used to maintain isothermal conditions and Peltier elements were


implemented to compensate the change of heat-transfer coefcient during measurements, making time-consuming calibration
unnecessary. The new combined reaction calorimeter has been
tested using three chemical reactions: neutralization of sodium
hydroxide with sulfuric acid, hydrolysis of acetic anhydride, and
acetylation of a substituted benzopyranol. The hydrolysis of acetic
anhydride was analyzed at three different temperature levels and
the total power produced during the reaction and the IR-ATR
measurements were used to quantify rate constant and activation
energy, respectively. Because the reaction is carried out in quite
dilute 0.1 M HCl the reactions kinetics was assumed to be pseudorst-order. Haji and Erkey (2005) developed a reaction engineering experiment that employs in-situ Fourier transfer infrared
(FTIR) spectroscopy for monitoring concentrations of hydrolysis of
acetic anhydride carried out in a batch reactor and isothermally at
room temperature. Since no sampling is required, the analytical
technique developed allows the reaction kinetics to be observed
without disturbing the reaction mixture. Concentrations of acetic
anhydride and acetic acid were measured as a function of time,
and the reaction order and rate constant were determined by the
integral method of analysis. As the reactions were carried
out in the presence of excess water, its concentration was not
monitored and the reaction kinetics was assumed to be pseudorst-order. Ampelli et al. (2003) used an isoperibolic reaction
calorimeter combined with an UVvis absorption spectrometer to
study the kinetics of hydrolysis of acetic anhydride. An acidbase
indicator is added to the reaction mixture and the change of its
color as a function of the extent of the hydrolysis reaction
is followed by visible spectrum measurements (optical pH
measurement).
Kralj (2007) determined the kinetics parameters (activation
energy, reaction rate constant, rate order) of hydrolysis of acetic
anhydride by measuring the conductivity of the weak electrolyte
(acetic acid). The reactions were carried out in a stirred batch
reactor at three different temperatures and using the molar ratio
of acetic anhydride to water equal to 1:131. The acetic acid
concentration was calculated on the theoretical basis of weak
electrolytes ionization, and the results obtained by the authors
showed that the hydrolysis of acetic anhydride is a pseudo-rstorder reaction.
In view of the foregoing, it can be seem that most classical
methods for the determination of kinetic parameters of hydrolysis
of acetic anhydride have relied mainly on reactors operated
isothermally at several pre-selected temperatures, and on a
variety of methods, some very difcult and time consuming, to
measure concentrations at discrete time intervals and that can be
severely limited when the reaction rates are fairly fast. Purely
calorimetric approaches do not require sampling and chemical
analysis, but may involve different levels of sophistication in the
equipment (reaction calorimeter) and in the data treatment.
Under certain conditions, calorimetric methods allow for determinination of temperature dependency of the rate constant in a
single experiment.
The aim of the present work is to explore the usefulness of
calorimetry as a tool to obtain information about liquid-phase
kinetics during the hydrolysis of acetic anhydride using temperature as the only measured variable in a very simple experimental
setup. Temperature versus time curves were used to determine
the frequency factor (k0), activation energy (E), and the global heat
exchanged (UA) for three reactions carried out under nonisothermal and non-adiabatic conditions (isoperibolic conditions).
The predicted conversion and reactor temperatures were also
determined for each reaction. The hydrolysis of acetic anhydride
was chosen as a test reaction because it is simple and several
literature references are available.

3851

2. Methodology
In the present work the kinetics of hydrolysis of acetic
anhydride was studied using simple isoperibolic calorimetry,
under non-isothermal conditions. The experiment is very simple,
employing only an arbitrary vessel (e.g., laboratory glassware)
and a digital thermocouple for measuring the variation of
temperature. Data treatment involves the use of mass and energy
balances to extract information regarding the kinetic parameters
of the reaction. As the reactor is not adiabatic, it is necessary to
include in the data treatment the determination of heat transfer
coefcient between the reactor and the surroundings .
2.1. Mass and energy balances
For a homogeneous reaction carried out in a liquid-phase,
constant-volume, non-adiabatic batch reactor, the mass and
energy balance equations, taking acetic anhydride (species A) as
the limiting reactant, are given by
NA0

dXA
rA V
dt

dTr
rA VDHUATr Tamb
6
dt
It is important to note that the value of (mCp)r in Eq. (6)
accounts for the heat capacity of the chemical components, as
well as for the contribution from heat capacities of the reactor
wall, stirrer, and components:
X
mj Cp,j
7
mCp r mCp reactor

mCp r

The effect of stirring was neglected in the energy balance.


Previous experiments without reaction (only water inside the
reactor, agitation turned on, and temperature measured for times
longer than the typical duration of the experiment with reaction)
have shown no detectable changes in temperature, so that the
heat generated by stirring was considered negligible.
The hydrolysis of acetic anhydride follows second-order
kinetics (rst order with respect to each reactant):
rA kCA CW

where k is the rate constant, which is temperature-dependent


according to the Arrhenius equation:
k k0 eE=RTr

In Eq. (8), the molar concentrations of acetic anhydride (CA)


and water (CW) may be written in terms of the conversion of the
limiting reactant as follows:
NA0
1XA
V


NA0 NW0

XA
V
NA0

CA

10

CW

11

Besides the initial amounts of water and acetic anhydride (NA0,


NW0), the room temperature (Tamb) and the temperature inside the
reaction (Tr(t)) are the only measured variables during the
experiment. The calculation procedure described below allows
one to estimate the evaluation of conversion (XA(t)) and the
kinetic (ko and E) and heat transfer (U) parameters.
For other reactions also studied by reaction calorimetry, such
as polymerizations, the heat transfer coefcient U could change
due to an increase in the viscosity of the reaction medium (e.g.,
Poc- o et al., 2010). In general, changes in viscosity during the
reaction should be a concern in other reactions (e.g., polymerization reactions), because the heat transfer coefcient varies with
liquid viscosity. However, in the reaction under study (hydrolysis

ARTICLE IN PRESS
3852

W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

of acetic anhydride) the changes in viscosity are not so intense;


therefore a constant value of the heat transfer coefcient can be
used throughout the reaction.
2.2. Determining the kinetic parameters, conversion and global heat
coefcient from the energy and mass balances for a non-adiabatic
system
In a non-adiabatic reactor, changes in temperature are related
to the rate of heat release by the exothermic reaction and to heat
exchange between the reactor and surroundings. After complete
conversion of the limiting reactant (acetic anhydride), temperature decreases due to heat exchange between the reactor and
surroundings. In the simple experimental setup used here, the
reactor is allowed to cool by heat exchange to air at ambient
temperature (this is called isoperibolic calorimetry).
Hence, the global heat transfer coefcient can be obtained
from an analysis of the transient during the temperature decrease
because the change of temperature occurs due to thermal
exchange between the reactor and the environment, since all
limiting reactant has been consumed. For this temperature
decrease, the governing equation is
mCp r

dTr
UATr Tamb
dt

12

Using the integrated form of Eq. (12), the slope of a plot of


ln(Tr  Tamb) versus t corresponds to UA/(mCp)r, from which the
global heat transfer coefcient can be determined.
Thus, given UA and the known parameters (  DH) and (mCp)r,
the conversion of acetic anhydride can also be estimated by
combining the mass and energy balances and integrating the
resulting equation from t0 to tf assuming that the conversion of
limiting reactant is equal to 100% at the end of the experiment.
Therefore, the experimental values of conversion of acetic
anhydride are given by the following recursive relationship:
Rt
mCp r Tr,i 1 Tr,i UA tii 1 Tr Tamb dt
13
XAi XAi 1 
DHNA0
Table 1
Room temperature.

rA

mCp r dTr =dt UATr Tamb


VDH

14

Once (  rA) is determined, the values of the rate constant at


each time can be calculated from Eq. (8) as
k

rA
CA CW

15

where CA and CW are calculated from Eqs. (10) and (11),


respectively.
Finally, the values of the frequency factor (k0) and the
activation energy (E) are obtained, respectively, from the intercept
and the slope of the straight line adjusted to the plot of ln(k)
versus 1/Tr:
lnk lnk0 

E1
R Tr

16

In Eq. (14), the values of the derivative dTr/dt are calculated by


nite difference method, using the following 2nd order central
nite difference or three-point differentiation formulas:
initial point

dTr 
3Tr t1 4Tr t2 Tr t3

dt 
2Dt

17

interior points


Tr tj 1 Tr tj1
dTr 

2Dt
dt 
t

18

last point

dTr 
Tr tn2 4Tr tn1 Tr tn

dt t
2Dt

19

t0

Reactor

Room temperature Tamb (K)

1
2
3

294.65
295.15
294.45

3. Experimental
The hydrolysis of acetic anhydride was carried out with excess of
water and in three different non-adiabatic vessels: (a) a cylindrical
plastic vessel (reactor 1), (b) a volumetric ask (reactor 2), and (c) a
thermal bottle or Dewar ask (reactor 3). Throughout the whole

Table 2
Reactants and reactors: volume and masses.
Volume (mL)
Deionized water
Acetic anhydride
Cylindrical plastic vessel
Volumetric ask
Thermal bottle
Magnetic stirrer

The integral of Eq. (13) was calculated for each interval of time
using the trapezoidal rule. Hence, the values of experimental
conversion were obtained from the nal time for complete
conversion of A backward to the starting time.
Knowing the value of UA, the rate of reaction of the limiting
reactant can be determined experimentally from the thermal
balance as follows:

Mass (g)

150
90

6.68
80.10
23.59
4.40

Table 4
Specic heat capacity of materials of reactors and stirrer.

Cylindrical plastic vessel


Volumetric ask
Thermal bottle
Magnetic stirrer

Cp (J/g K)

Ref.

1.79
0.75
0.75
0.65

Ullmann (1985)
Ullmann (1985)
Ullmann (1985)
Westrum and Grnvold (1969)

Table 3
Constants for specic heat capacity of reactants and products.
A (J/mol K)
Water
Acetic anhydride
Acetic acid

92.053
71.831
 18.944

B (J/mol K2)

C (J/mol K3)
2

 3.995  10
8.888  10  1
1.0971

D (J/mol K4)
4

 2.211  10
 2.653  10  3
 2.892  10  3

7

5.347  10
3.350  10  6
2.93  10  6

Ref.
Yaws (1999)
Yaws (1999)
Yaws (1999)

ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

reaction, the reaction medium was kept well stirred by a magnetic


stirrer, and the reactor surrounds were at room temperature. As the
reactions were carried out on different days, the room temperatures
were different for each type of reactor and are reported in Table 1.

3853

Table 6
Initial conversions.
Reactor

XA at t 0 (%)

1
2
3

6.33
4.94
5.72

Fig. 1. Graphical estimation of global heat transfer for the cooling region for:
(a) reactor 1, (b) reactor 2, and (c) reactor 3.

Table 5
Experimental heat transfer coefcient.
Reactor

UA (W/K)

1
2
3

0.424
0.280
0.0846

Fig. 2. Arrhenius plot for the data taken from: (a) reactor 1, (b) reactor 2, and
(c) reactor 3.

ARTICLE IN PRESS
3854

W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

As already mentioned, the heating effect of stirring was evaluated in


experiments with water (no reaction) for longer periods and no
detectable change in temperature was measured. Thus, the heating
effect of stirring was neglected.
The rst two systems were used to simulate a non-adiabatic
reactor while reactor 3 simulated a non-ideal adiabatic reactor.
These systems were chosen to allow the implementation of this
study as an experiment for the undergraduate laboratory. The
temperatures were recorded every 30 s using a digital thermometer with a sheathed thermocouple.
Acetic anhydride having a minimum assay of 95% and
deionized water were used and the volumes of both are shown
in Table 2 along with masses of the vessels and the magnetic
stirrer (magnetiteFe3O4).
The specic heat capacity of the reactants and products was
calculated by
Cp A BT CT 2 DT 3

20

The constants AD are presented in Table 3. Specic heat


capacities of the reactor wall and the magnetic stirrer are
presented in Table 4 and the heat of reaction for the hydrolysis
of acetic anhydride was assumed to be (  DH) 58,994.4 J/mol
(Conn et al., 1942).
In Table 1 the mass of the thermal bottle was calculated
disregarding the external wrapper and considering that the mass
of glass ampoule corresponds to 10% of the total mass of the
bottle. The external wrapper was not considered because we
consider that only the internal wall of the ampoule reaches the
same temperature of the liquid inside the reactor. The specic
heat capacities of the volumetric ask and glass ampoule (Table 3)
are related to the specic heat capacity of an alloy composed 96%
of silicon oxide.

Table 7
Experimental activation energy and ln(k0) values for the hydrolysis of acetic
anhydride.
Reactor

ln(k0) (L/mol s)

E (kJ/mol)

1
2
3

16.25
15.28
15.15

68.9
66.5
66.3

4. Results and discussion


In order to determine experimental values of activation energy
and frequency factor, the global heat transfer coefcient must be
known. This coefcient can be experimentally inferred from
analysis of temperature variations during the temperature
decrease where no reaction occurs due to complete conversion
of the limiting reactant. Taking the integrated form of Eq. (12), UA
can be readily estimated from the slope of a plot of ln(TrTamb)
versus t (Fig. 1). The global heat capacity of the reactor, used in
Eq. (12), was computed by considering the heat capacity of the
reactor and the magnetic stirrer plus the average heat capacities of
water and acetic acid produced by complete conversion of acetic
anhydride. The estimated values of UA are presented in Table 5.
The correlation coefcients in Fig. 1 ranged from 0.997 to 0.999.
Comparing the estimated values of heat transfer coefcients
presented in Table 5, it can be seen that the value of UA for the
volumetric ask is slightly lower than that obtained for reactor 1.
This occurs because the heat transfer resistance of reactor 2 is
greater than that of the polystyrene cup. Besides, the long neck of
the volumetric ask makes heat exchange between the reaction
medium and the environment difcult.
It is important to note that while heat exchange through the
double walls of the thermal bottle is minimal, reactor 3 presents a
nonzero value of the heat transfer coefcient, because the thermal
bottle was kept open throughout the reaction, allowing for some
heat exchange though the open top of the bottle.
Using the estimated values of UA, the experimental prole for
conversion can be obtained by recursive application of Eq. (13).
Because all reactions were carried out with an excess of deionized
water, conversions were computed from the nal point, where the
conversion of acetic anhydride was assumed to be equal to 100%.
Values of the initial conversion cannot be assumed to be zero
because the purity of the acetic anhydride is not well known.
Table 6 lists the initial conversion for each system. It can be
seen that the experimental conversions at t 0 are very different
from zero. Therefore, apparently the actual purity of the acetic
anhydride used during the reactions is slightly smaller than the
nominal purity provided by manufacturer. The decrease in purity
may be caused by the duration of storage of the reagent or contact
with the environment since acetic anhydride reacts with moisture
in air.
Regarding the kinetic parameters, using the experimental
values of temperature, global heat transfer coefcient (Table 5),

Table 8
Kinetic parameters for hydrolysis of acetic anhydride reported in the literature.
Ref.

Acetic anhydride concentration


(mol/L)

Measuring techniques

E (kJ/mol)

King and Glasser (1965)


Eldridge and Piret (1950)
Takashima et al. (1971)
Cleland and Wilhelm (1956)
Dyne et al. (1967)
Bisio and Kabel (1985)
Glasser and Williams (1971)
Takashima et al. (1971)
Shatynski and Hanesian (1993)
Wilsdon and Sidgwick (1913)
Kralj (2007)
Rivett and Sidgwick (1910)
Haji and Erkey (2005)
Asprey et al. (1996)
Wilsdon and Sidgwick (1913)
Marek (1954)
Marmers (1965)

n.r.
n.r.
Innite dilution
0.020.06
0.17
0.22
0.25
0.25
0.27
0.34
0.41
0.54
0.66
1.0
1.10
1.34
Equimolar concentration

Calorimetry
Titration
n.r.
Titration
Calorimetry
Calorimetry
Calorimetry
n.r.
Calorimetry
Conductivity
Conductivity
Conductivity
FTIR
Conductivity
Conductivity
n.r.
n.r

39.8
43.2
33.6
44.4
49.4
46.5
45.3
49.4
46.9
50.6
50.1
43.2
53.6
45.7
56.2
57.8
68.7

n.r. not reported.

ln(k0)

Units of k0

7.80

s1
L/(mol s)

L/(mol s)

12.80
7.95
18.1
12.74
18.52
14.21

s1
L/(mol s)
s1
s1
s1
s1

15.48
7.66
20.45

s1
L/(mol s)
s1

9.93
7.53

ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

and conversion, both k0 and E can be easily computed graphically


via Eq. (16). Fig. 2 shows a typical Arrhenius plot according to
Eq. (16), where the slope is equal to (  E/R) and the intercept (at
1/Tr 0) is ln(k0). As can be seen from Fig. 2, the data fall on a
reasonably straight line with correlation coefcients above 0.99.
The kinetic parameters (k0 and E) obtained for each system are
summarized in Table 7. The literature values for the activation
energy and frequency factor for the hydrolysis of acetic anhydride
are listed in Table 8, based on the values reported by Asprey et al.
(1996) and Shatynski and Hanesian (1993) and complemented

3855

with additional literature information. It is important to note that,


in Table 8, some frequency factors were obtained under the
hypothesis of pseudo-rst-order reaction conditions. Therefore,
for comparison purposes, these parameters should be converted
to second-order constants by dividing the pseudo-rst-order
constant by the concentration of water, when available.
Comparing the kinetic parameters presented in Table 7
obtained in the present work, it can be observed that the values
are very similar, regardless of the reactor used. However, when
the values of Table 7 are compared with those found in the
literature (Table 8), it can be seen that the values of activation
energy and frequency factor obtained are always higher than
those reported in Table 8. It is well known that small differences
in activation energy causes large differences in ln(k0), as this value
is obtained by extrapolation of the Arrhenius plot far from the
studied range of temperature. Therefore, it is sounder to compare
directly the values of k obtained at different temperatures, as
shown in Fig. 3. The obtained values of k are within the range of
literature values for higher temperatures, and larger deviations
are found for the low temperature range. A possible explanation
for these deviations is that, for low temperatures, solubility of
acetic anhydride in water is not complete. The excess of acetic
anhydride is rst dispersed as droplets, and then completely
dissolves over time, favored by the increase of temperature and
the formation of some acetic acid from the reaction. Indeed, in the
experiments in glass reactor 2, it was possible to see that the
reaction mixture was initially turbid due to the presence of small
droplets, a clear indication of a partially immiscible system (if the
stirring was stopped, droplet coalescence and phase separation
could be observed). When the temperature is higher than about
35 1C, the appearance of the mixture suddenly changes to a clear
solution, indicating a completely miscible system from this point
on. Thus, at rst only the data taken after this full solubilization
point should be considered.

Table 9
Experimental activation energy and ln(k0) values for the hydrolysis of acetic
anhydride without mixing effects.
Reactor

ln(k0) (L/mol s)

E (kJ/mol)

1
2
3

17.57
16.65
16.19

73.74
70.14
69.12

100

E (kJ/mol)

80

60

literature data

40

present work
20

present work, considering


only the data for T > 35 C

0
0

Fig. 3. Arrhenius plot for comparing the rate constant for the hydrolysis of acetic
anhydride obtained in this work with those reported in the literature: (a) reactor 1,
(b) reactor 2, and (c) reactor 3.

2
4
6
8
initial concentration of acetic anhydride (mol/L)

10

Fig. 4. Effect of initial concentration of acetic anhydride on activation energy of


the hydrolysis reaction. Each point corresponds to one work from the literature,
according to Table 8.

ARTICLE IN PRESS
3856

W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

In order to eliminate the above mentioned effect of incomplete


mixture during the beginning of the experiment, the kinetic
parameters were also computed using the temperatures measured only after the initial disturbances disappeared (Tr 435 1C).
The new kinetic parameters thus estimated are summarized in
Table 9.
Comparing the values of activation energy and ln(k0) presented in Tables 7 and 9, it can be seen that the results are very
similar, implying that solubility is not mainly responsible for the
variability in reaction rates. In fact, Golding and Dussault (1978)
noted that data in the literature indicated that in the case of
excess of either reactant there is deviation from second order

kinetics, and both frequency factor and activation energy increase


with increasing acetic anhydride concentration. Consequently, for
the anhydride concentration range studied in the literature
(0.2911.47 mol/L) no unique value of either ln(k0) or E could be
taken. Besides, Janssen et al. (1957) noted a decrease in the rate
constant in the presence of concentrated acetic acid. This fact can
be attributed to the formation of hydrates of acetic acid, as quoted
by Plyler and Barr (1935). According to these authors, the
hydration of some of the water with acetic acid formed probably
keeps the water from having a part in the reaction and the
presence of acetic acid also decreases the number of collisions in a
given time between the anhydride and water molecules. On the

Fig. 5. Experimental and predicted values of temperature for: (a) reactor 1,


(b) reactor 2, and (c) reactor 3.

Fig. 6. Experimental and predicted values of conversion for: (a) reactor 1,


(b) reactor 2, and (c) reactor 3.

ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

other hand, Golding and Dussault (1978) also observed that the
effect of acetic acid on reaction rate reported in the literature has
been conicting.
Fig. 4 shows the variation of activation energy obtained by
different authors, as a function of the initial concentration of
acetic anhydride. It is quite evident that, for smaller
concentrations, activation energy increases with increasing
concentration of acetic anhydride. Our results, obtained for
relatively higher initial concentration of acetic anhydride, are
also included in this plot and follow reasonably well the trends of
changes of E with temperature.
Finally, in order to test the values obtained for the 3
parameters (U, k0, and E), they were used in the energy and mass
balances to simulate the evaluation of temperature and conversion. Figs. 5 and 6 show a comparison between the experimental
values of the reactor temperature (Fig. 5) and the conversion
(Fig. 6) with those predicted by the energy and mass balances
using the estimated values of UA and kinetic parameters (k0 and E)
for each case analyzed. For all cases, a good agreement between
experimental and predicted values was found despite values of
activation energy and frequency factor obtained being always
higher than those reported in literature.

5. Conclusions
This work shows that both kinetic parameters and heat
transfer coefcient of reactor contents can be easily determined
for hydrolysis of acetic anhydride using temperature-versus-time
data under non-adiabatic, isoperibolic conditions. The main
difference of this work in relation to other previously published
works in the literature is that the kinetic data were obtained
under non-adiabatic conditions, where the heat transfer coefcient must be known in order to determine these parameters. The
results shows that the estimates obtained for activation energy
and frequency factor of the hydrolysis of acetic anhydride are
apparently affected by excess of either reactant and, therefore, no
unique value of either ln(k0) or E could be taken. Despite the
observed differences between estimates of k0 and E and those
reported in the literature, experimental and predicted values of
temperature and conversion show good agreement.
As a nal comment, the simplicity of the experimental setup
and the calculations used in this investigation allows one to use
them as an experiment for undergraduate laboratories to
illustrate the application of the non-adiabatic calorimetric
method to infer state variables.

Notation
CA
CW
Cp
E
k
k0
(mCp)r
MW
NA0
NW0
R
( rA)
t0
tf
Tamb
Tr
UA

concentration of acetic anhydride, mol/L


concentration of water, mol/L
specic heat capacity, J/mol K or J/g K
activation energy, J/mol
specic reaction rate constant, L/mol s
pre-exponential factor or frequency factor, L/mol s
total heat capacity of reactor and contents, J/K
molecular weight, g/mol
moles of acetic anhydride initially fed to reactor, mol
moles of water initially fed to reactor, mol
gas constant, J/mol K
rate of disappearance of component A, mol/L s
initial instant, s
nal instant, s
environment temperature, K
reactor temperature, K
heat transfer coefcient, J/K s

3857

V
total volume of reactants added to reactor, L
XA
conversion of component A
(  DH) heat of hydrolysis of acetic anhydride, J/mol
r
density, g/L

Acknowledgements
The nancial supports from FAPESPFundac- a~ o de Amparo a
Pesquisa do Estado de Sa~ o Paulo, CNPqConselho Nacional
de Desenvolvimento Cientco e Tecnologico, and CAPES
Coordenac- a~ o de Aperfeic-oamento de Pessoal de Nvel Superior
are gratefully appreciated.
References
Ampelli, C., Di Bella, D., Lister, D.G., Maschio, G., Parisi, J., 2003. The integration of
an ultravioletvisible spectrometer and a reaction calorimeter. Journal of
Thermal Analysis and Calorimetry 72, 875883.
Ampelli, C., Di Bella, D., Lister, D.G., Maschio, G., 2005. Fitting isoperibolic
calorimeter data for reactions with pseudo-rst order chemical kinetics.
Journal of Thermal Analysis and Calorimetry 79, 8994.
Asprey, S.P., Wojciechowski, B.W., Rice, N.M., Dorcas, A., 1996. Applications of
temperature scanning in kinetic investigations: the hydrolysis of acetic
anhydride. Chemical Engineering Science 51 (20), 46814692.
Bell, W.C., Booksh, K.S., Myrick, M.L., 1998. Monitoring anhydride and acid
conversion in supercritical/hydrothermal water by in situ ber-optic Raman
spectroscopy. Analytical Chemistry 70 (2), 332339.
Bisio, A., Kabel, R.L., 1985. Scaleup of Chemical Process. Wiley, New York (pp. 136138).
Cleland, F.A., Wilhelm, R.H., 1956. Diffusion and reaction in viscous-ow tubular
reactor. AIChE Journal 2, 489.
Conn, J.B., Kistiakowsky, G.B., Roberts, R.M., Smith, E.A., 1942. Heats of organic
reactions. XIII. Heats of hydrolysis of some acid anhydrides. Journal of the
American Chemical Society 64 (8), 17471752.
Dyne, S.R., Glasser, D., King, R.P., 1967. Automatically controlled adiabatic reactor
for reaction rate studies. The Review of Scientic Instruments 38 (2), 209214.
Eldridge, J.W., Piret, E.L., 1950. Continuous-ow stirred-tank reactor system I.
Design equations for homogeneous liquid phase reactions. Experimental data.
Chemical Engineering Progress 46, 290.
Glasser, D., Williams, D.F., 1971. The study of liquid-phase kinetics using
temperature as a measured variable. Industrial and Engineering Chemistry
Fundamentals 10 (3), 516519.
Gold, V., 1948. The hydrolysis of acetic anhydride. Transactions of the Faraday
Society 44, 506518.
Golding, J.A., Dussault, R., 1978. Conversions and temperature rises in a laminar
ow reactor for the hydration of acetic anhydride. The Canadian Journal of
Chemical Engineering 56, 564569.
Haji, S., Erkey, C., 2005. Kinetics of hydrolysis of acetic anhydride by in-situ FTIR
spectroscopy. An experiment for the undergraduate laboratory. Chemical
Engineering Education 39 (1), 5661.
Janssen, H.J., Haydel, C.H., Greathouse, L.H., 1957. Hydrolysis of acetic anhydride in
concentrated acetic acid without catalysis. Industrial and Engineering
Chemistry 49 (2), 197201.
King, R.P., Glasser, D., 1965. The use of the adiabatic calorimeter for reaction rate
studies. The South African Industrial Chemistry 19, 1215.
Kralj, A.K., 2007. Checking the kinetics of acetic acid production by measuring the
conductivity. Journal of Industrial and Engineering Chemistry 13 (4), 631636.
Marek, J., 1954. Collection of Czechoslovak Chemical Communications 19, 621.
Marmers, H., 1965. Ph.D. Dissertation, University of Birmingham, Birmingham,
England.
Oakenfull, D.G., 1974. The kinetics of the hydrolysis of acetic anhydride and the
reaction of 4-nitrophenyl acetate with imidazole in aqueousorganic mixed
solvents. Australian Journal of Chemistry 27 (7), 14231431.
Orton, K.J.P., Jones, M., 1912. Hydrolysis of acetic anhydride. Journal of the
Chemical Society 101, 17081720.
Plyler, E.K., Barr, E.S., 1935. The reaction rate of acetic anhydride and water. Journal
of Chemical Physics 3 (11), 679682.
Poc- o, J.G.R., Danese, M., Giudici, R., 2010. Investigation of cationic polymerization
of beta-pinene using calorimetric measurements. Macromolecular Reaction
Engineering 4 (2), 145154.
Puxty, G., Maeder, M., Rhinehart, R.R., Alam, S., Moore, S., Gemperline, P.J., 2005.
Modeling batch reactions with in situ spectroscopy measurements and
calorimetry. Journal of Chemometrics 19, 329340.
Regenass, W., 1985. Calorimetric monitoring of industrial chemical process.
Thermochimica Acta 95, 351368.
Rivett, A.C.D., Sidgwick, N.V., 1910. The rate of hydration of acetic anhydride.
Journal of the Chemical Society 97, 732741.
Shatynski, J.J., Hanesian, D., 1993. Adiabatic kinetic studies of the cytidine/acetic
anhydride reaction by utilizing temperature versus time data. Industrial and
Engineering Chemistry Research 32 (4), 594599.

ARTICLE IN PRESS
3858

W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

Smith, T.L., 1955. Application of ice calorimetry to chemical kinetics. Journal of


Physical Chemistry 59 (5), 385388.
Takashima, I., Nishida, A., Ogita, K., Uchida, N., 1971. Development of liquid phase
reaction analyser 3. Computer aided reaction - analysis on twin calorimetry.
Kogyo Kagaku Zasshi 74, 12931297.
Ullmann, F., 1985. Ullmanns encyclopedia of industrial chemistry, 5th ed WileyVCH, Weinheim.

Visentin, F., Gianoli, S.I., Zogg, A., Kut, O.M., Hungerbuhler,


K., 2004. A pressureresistant small-scale reaction calorimeter that combines the principles of
power compensation and heat balance. Organic Process Research and
Development 8 (5), 725737.
Westrum, E.F., Grnvold, F., 1969. Magnetite (Fe3O4) heat capacity and thermodynamic properties from 5 to 350 K, low-temperature transition. The Journal
of Chemical Thermodynamics 1 (6), 543557.

Wilsdon, B.H., Sidgwick, N.V., 1913. The rate of hydration of acid anhydrides: acetic,
propionic, butyric, and benzoic. Journal of the Chemical Society 103, 19591973.
Wojciechowski, B.W., 1997. The temperature scanning reactor I: reactor types and
modes of operation. Catalysis Today 36, 167190.
Yaws, C.L., 1999. Chemical Properties Handbook: Physical, Thermodynamic,
Environmental, Transport, Safety, and Health Related Properties for Organic
and Inorganic Chemicals. McGraw Hill, New York.

Zogg, A., Fischer, U., Hungerbuhler,


K., 2003. A new small-scale reaction
calorimeter that combines the principles of power compensation and heat
balance. Industrial and Engineering Chemistry Research 42 (4), 767776.

Zogg, A., Fischer, U., Hungerbuhler,


K., 2004. A new approach for a combined
evaluation of calorimetric and online infrared data to identify kinetic and
thermodynamic parameters of a chemical reaction. Chemometrics and
Intelligent Laboratory Systems 71 (2), 165176.

Potrebbero piacerti anche