You are on page 1of 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
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
b
Universidade Federal de Santa Catarina, Departamento de Engenharia Qumica e de Alimentos, Floriano polis, Brazil

a r t i c l e in fo abstract

Article history: A simple calorimetric method to estimate both kinetics and heat transfer coefcients using
Received 14 August 2009 temperature-versus-time data under non-adiabatic conditions is described for the reaction of
Received in revised form hydrolysis of acetic anhydride. The methodology is applied to three simple laboratory-scale reactors
14 March 2010
in a very simple experimental setup that can be easily implemented. The quality of the experimental
Accepted 19 March 2010
results was veried by comparing them with literature values and with predicted values obtained by
Available online 24 March 2010
energy balance. The comparison shows that the experimental kinetic parameters do not agree exactly
Keywords: with those reported in the literature, but provide a good agreement between predicted and
Acetic anhydride experimental data of temperature and conversion. The differences observed between the activation
Reaction calorimetry
energy obtained and the values reported in the literature can be ascribed to differences in anhydride-to-
Hydrolysis
water ratios (anhydride concentrations).
Kinetics
Chemical reactors & 2010 Elsevier Ltd. All rights reserved.
Reaction engineering
Mathematical modeling

1. Introduction anhydride can be represented as


O
During the early stages of a chemical process development,
H3C C H3C OH
tasks such as delineating safety conditions and reaction optimiza-
tion and control can be carried out only if a reaction model and C
O + H2O 2 1
the corresponding reaction parameters are known. If the kinetics
H3C C
is not fully understood, several problems may arise, including O
runaway reactions that can lead to the release of environmentally O
dangerous substances, serious physical equipment damage, or
The overall reaction mechanism proceeds via three irreversible
even the possibility of an explosion. Unfortunately, for the
steps: addition, elimination, and proton transfer to solvent
majority of reactions, there is hardly any knowledge about the
(Asprey et al., 1996):
kinetic, thermodynamic, and physical parameters.
addition
Since several chemical reactions are moderatelyhighly
exothermic, it is possible to quantify continuously the amount CH3
of heat released based only on temperature measurements and O O
CH3 CH3
energy balance equations that, in turn, can be used to infer useful
information about the progress of the reaction such as thermo- C C + H2O CH3 C C 2
+
chemical and kinetics parameters, reaction calorimetry becoming O O O O O
an essential tool for data-oriented development of chemical H
H
engineering processes.
Hydrolysis of acetic anhydride is a moderatelyhighly exothermic, elimination
fast reaction, which is ideal for verifying the dynamic response of CH3
a calorimetric reactor. The overall hydrolysis reaction of acetic O
O O O CH3

CH3 C C CH3 C + C 3
+ +
 Corresponding author. Fax: +55 11 3813 2380. O
O O
E-mail address: rgiudici@usp.br (R. Giudici). O
1 H H H H
Fax: +55 48 3331 9687.

0009-2509/$ - see front matter & 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2010.03.028
ARTICLE IN PRESS
3850 W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858

proton transfer 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
4
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
where the addition reaction is the rate-controlling process, and noise in the conductivity signal was unavoidably present due to
the resulting kinetics are rst-order in each of the reactants the electrodes used. The operation of a temperature scanning
(Asprey et al., 1996). reactor involves the ramping of the feed temperature while
The kinetics of the hydrolysis of acetic anhydride have been collecting composition and temperature data at the output of the
studied in literature using different techniques for measuring the reactor. In order to make such data interpretable, the temperature
extent of the reaction and the reaction rate: titration (Orton and ramping rate must be such that the time between successive
Jones, 1912; Cleland and Wilhelm, 1956), colorimetry (Oakenfull, analyses is much shorter than the time required for a kinetically
1974), conductivity (Rivett and Sidgwick, 1910; Asprey et al., signicant increment of the temperature to be induced in the feed
1996; Kralj, 2007), spectroscopic techniques (Bell et al., 1998; to the reactor (Wojciechowski, 1997). As the reactions were
Zogg et al., 2004; Haji and Erkey, 2005), and different calorimetric carried out with an excess of water, pseudo-rst-order reaction
techniques (Gold, 1948; Smith, 1955; Janssen et al., 1957; Glasser conditions were assumed. Comparison of Arrhenius plot deter-
and Williams, 1971; Regenass, 1985; Shatynski and Hanesian, mined by temperature scanning with those found in the literature
1993; Ampelli et al., 2003, 2005). More recently, the combination showed that the reaction constant obtained by this method is
of calorimetric and spectroscopic techniques was also proposed smaller than those obtained by the methods previously used in
(Zogg et al., 2003, 2004; Puxty et al., 2005). studies of acetic anhydride hydrolysis.
Conn et al. (1942) employed calorimetry to measure the Bell et al. (1998) monitored the hydrolysis of acetic anhydride
enthalpy of hydrolytic reactions at 303 K for a number of straight- in a hydrothermal/supercritical water reactor with Raman
chain and cyclic acid anhydrides (acetic, propionic, isobutyric, and spectroscopy to demonstrate the use of in situ spectroscopy
trimethylacetic anhydrides). Gold (1948) analyzed the effect of combined with mathematical interpretation to estimate kinetics
solvent type (water and acetonewater mixtures) and tempera- and thermodynamics parameters of an organic reaction carried
ture on the kinetics of the hydrolysis of acetic anhydride; rst- out in a challenging medium. From the intensity relationships and
order rate constants were determined by conductimetry at 5, 15, the temperature information, the activation energy of the
and 25 1C, and the data analyzed in terms of the parameters of the hydrolysis reaction was estimated and compared to literature
Arrhenius equation; the reaction belongs to the class of values. Besides, the authors also estimated the enthalpic and
nucleophilic bimolecular (SN2) solvolytic substitution reactions entropic cost of this reaction for a change in solvation environ-
and, therefore, the hydrolysis of acetic anhydride obeys second ment of acetic acid at high temperatures. The authors noted that
order kinetics, being rst order with respect to both acetic the value of activation energy obtained by this method is
anhydride and water. In order to elucidate the role of water consistent with the literature value of acid-catalyzed hydrolysis,
structure in the kinetics of hydrolysis, Oakenfull (1974) made a but signicantly lower than those values reported for noncata-
comparative study of the kinetics of hydrolysis of acetic lyzed process. The experiments were carried out in a stainless
anhydride and the reaction of 4-nitrophenyl acetate with steel high-pressure reactor vessel with internal volume of 10.4 mL
imidazole in mixtures of water with ethanol, t-butyl alcohol, using 2 mL of distilled water and 2 mL of acetic anhydride.
dimethyl sulphoxide, and dioxin; he observed that both rate Zogg et al. (2004) proposed a new approach that combines the
constants were always reduced by the addition of organic solvent. joint evaluation of calorimetric and online infrared data to
Janssen et al. (1957) used calorimetric techniques to study the identify kinetic and thermodynamic parameters of a chemical
hydrolysis of acetic anhydride in concentrated acetic acid without reaction. Furthermore, a new weighting principle was developed
catalyst. that performs an automatic scaling of the infrared and calori-
Dyne et al. (1967) and King and Glasser (1965) measured the metric data in order to equalize their inuence on the estimated
kinetics of hydrolysis of acetic anhydride in diluted solutions reaction parameters. According to the authors, the evaluation of
using an adiabatic calorimetric reactor in which the reactor walls the calorimetric and infrared data is conventionally carried out
were electrically heated to be at the same temperature of the uid separately. This will generally lead to different estimated reaction
inside the reactor. Later, Glasser and Williams (1971) determined parameters caused by measurement errors, by unmodeled
the kinetic parameters of the hydrolysis of acetic anhydride in processes with different inuences on the two analytical signals,
dilute aqueous solution, from experimental temperaturetime or unequal information content of the two signals. The feasibility
curves by a regression analysis on the differential equations of the new evaluation algorithm was demonstrated based on the
describing the reaction system. A small vessel reactor, working as hydrolysis of acetic anhydride. The authors noted that the
an isoperibolic calorimeter, was placed in a constant-temperature reaction of acetic anhydride with water shows signicant heat
bath, and the heat transfer coefcient between reactor and the of mixing, this effect is visible only in the calorimetric signals and
bath was determined from temperaturetime curve obtained does not affect the infrared spectrum, and, therefore, the
using a known amount of a hot or cold suitable inert liquid added separated evaluation of the infrared and calorimetric data gave
in the reactor using a syringe lled with a known quantity of one similar results only when the dosing phase was excluded from the
reactant. In order to eliminate the mixing effects when the calorimetric data. As the reaction was carried out in acid medium,
reactants are rst brought together, the zero of the time scale of the reactions kinetics was assumed to be pseudo-rst-order. Zogg
differential equations that describe the reactions system was et al. (2003) and Visentin et al. (2004) developed a small-scale
chosen at an arbitrary point after the initial disturbances died reaction calorimeter tted with integrated infrared-attenuated
away. Shatynski and Hanesian (1993) also determined the total reection (IR-ATR), which combines the principles of power
kinetics of the hydrolysis reactions using temperature data compensation and heat balance. The power compensation was
ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858 3851

used to maintain isothermal conditions and Peltier elements were 2. Methodology


implemented to compensate the change of heat-transfer coef-
cient during measurements, making time-consuming calibration In the present work the kinetics of hydrolysis of acetic
unnecessary. The new combined reaction calorimeter has been anhydride was studied using simple isoperibolic calorimetry,
tested using three chemical reactions: neutralization of sodium under non-isothermal conditions. The experiment is very simple,
hydroxide with sulfuric acid, hydrolysis of acetic anhydride, and employing only an arbitrary vessel (e.g., laboratory glassware)
acetylation of a substituted benzopyranol. The hydrolysis of acetic and a digital thermocouple for measuring the variation of
anhydride was analyzed at three different temperature levels and temperature. Data treatment involves the use of mass and energy
the total power produced during the reaction and the IR-ATR balances to extract information regarding the kinetic parameters
measurements were used to quantify rate constant and activation of the reaction. As the reactor is not adiabatic, it is necessary to
energy, respectively. Because the reaction is carried out in quite include in the data treatment the determination of heat transfer
dilute 0.1 M HCl the reactions kinetics was assumed to be pseudo- coefcient between the reactor and the surroundings .
rst-order. Haji and Erkey (2005) developed a reaction engineer-
ing experiment that employs in-situ Fourier transfer infrared 2.1. Mass and energy balances
(FTIR) spectroscopy for monitoring concentrations of hydrolysis of
acetic anhydride carried out in a batch reactor and isothermally at For a homogeneous reaction carried out in a liquid-phase,
room temperature. Since no sampling is required, the analytical constant-volume, non-adiabatic batch reactor, the mass and
technique developed allows the reaction kinetics to be observed energy balance equations, taking acetic anhydride (species A) as
without disturbing the reaction mixture. Concentrations of acetic the limiting reactant, are given by
anhydride and acetic acid were measured as a function of time,
and the reaction order and rate constant were determined by the dXA
NA0 rA V 5
integral method of analysis. As the reactions were carried dt
out in the presence of excess water, its concentration was not dTr
monitored and the reaction kinetics was assumed to be pseudo- mCp r rA VDHUATr Tamb 6
dt
rst-order. Ampelli et al. (2003) used an isoperibolic reaction
It is important to note that the value of (mCp)r in Eq. (6)
calorimeter combined with an UVvis absorption spectrometer to
accounts for the heat capacity of the chemical components, as
study the kinetics of hydrolysis of acetic anhydride. An acidbase
well as for the contribution from heat capacities of the reactor
indicator is added to the reaction mixture and the change of its
wall, stirrer, and components:
color as a function of the extent of the hydrolysis reaction X
is followed by visible spectrum measurements (optical pH mCp r mCp reactor mj Cp,j 7
measurement). j

Kralj (2007) determined the kinetics parameters (activation The effect of stirring was neglected in the energy balance.
energy, reaction rate constant, rate order) of hydrolysis of acetic Previous experiments without reaction (only water inside the
anhydride by measuring the conductivity of the weak electrolyte reactor, agitation turned on, and temperature measured for times
(acetic acid). The reactions were carried out in a stirred batch longer than the typical duration of the experiment with reaction)
reactor at three different temperatures and using the molar ratio have shown no detectable changes in temperature, so that the
of acetic anhydride to water equal to 1:131. The acetic acid heat generated by stirring was considered negligible.
concentration was calculated on the theoretical basis of weak The hydrolysis of acetic anhydride follows second-order
electrolytes ionization, and the results obtained by the authors kinetics (rst order with respect to each reactant):
showed that the hydrolysis of acetic anhydride is a pseudo-rst-
rA kCA CW 8
order reaction.
In view of the foregoing, it can be seem that most classical where k is the rate constant, which is temperature-dependent
methods for the determination of kinetic parameters of hydrolysis according to the Arrhenius equation:
of acetic anhydride have relied mainly on reactors operated
k k0 eE=RTr 9
isothermally at several pre-selected temperatures, and on a
variety of methods, some very difcult and time consuming, to In Eq. (8), the molar concentrations of acetic anhydride (CA)
measure concentrations at discrete time intervals and that can be and water (CW) may be written in terms of the conversion of the
severely limited when the reaction rates are fairly fast. Purely limiting reactant as follows:
calorimetric approaches do not require sampling and chemical NA0
CA 1XA 10
analysis, but may involve different levels of sophistication in the V
equipment (reaction calorimeter) and in the data treatment.  
Under certain conditions, calorimetric methods allow for deter- NA0 NW0
CW XA 11
minination of temperature dependency of the rate constant in a V NA0
single experiment. Besides the initial amounts of water and acetic anhydride (NA0,
The aim of the present work is to explore the usefulness of NW0), the room temperature (Tamb) and the temperature inside the
calorimetry as a tool to obtain information about liquid-phase reaction (Tr(t)) are the only measured variables during the
kinetics during the hydrolysis of acetic anhydride using tempera- experiment. The calculation procedure described below allows
ture as the only measured variable in a very simple experimental one to estimate the evaluation of conversion (XA(t)) and the
setup. Temperature versus time curves were used to determine kinetic (ko and E) and heat transfer (U) parameters.
the frequency factor (k0), activation energy (E), and the global heat For other reactions also studied by reaction calorimetry, such
exchanged (UA) for three reactions carried out under non- as polymerizations, the heat transfer coefcient U could change
isothermal and non-adiabatic conditions (isoperibolic conditions). due to an increase in the viscosity of the reaction medium (e.g.,
The predicted conversion and reactor temperatures were also Poc- o et al., 2010). In general, changes in viscosity during the
determined for each reaction. The hydrolysis of acetic anhydride reaction should be a concern in other reactions (e.g., polymeriza-
was chosen as a test reaction because it is simple and several tion reactions), because the heat transfer coefcient varies with
literature references are available. 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; The integral of Eq. (13) was calculated for each interval of time
therefore a constant value of the heat transfer coefcient can be using the trapezoidal rule. Hence, the values of experimental
used throughout the reaction. conversion were obtained from the nal time for complete
conversion of A backward to the starting time.
2.2. Determining the kinetic parameters, conversion and global heat Knowing the value of UA, the rate of reaction of the limiting
coefcient from the energy and mass balances for a non-adiabatic reactant can be determined experimentally from the thermal
system balance as follows:
mCp r dTr =dt UATr Tamb
In a non-adiabatic reactor, changes in temperature are related rA 14
VDH
to the rate of heat release by the exothermic reaction and to heat
exchange between the reactor and surroundings. After complete Once (  rA) is determined, the values of the rate constant at
conversion of the limiting reactant (acetic anhydride), tempera- each time can be calculated from Eq. (8) as
ture decreases due to heat exchange between the reactor and rA
surroundings. In the simple experimental setup used here, the k 15
CA CW
reactor is allowed to cool by heat exchange to air at ambient
where CA and CW are calculated from Eqs. (10) and (11),
temperature (this is called isoperibolic calorimetry).
Hence, the global heat transfer coefcient can be obtained respectively.
Finally, the values of the frequency factor (k0) and the
from an analysis of the transient during the temperature decrease
activation energy (E) are obtained, respectively, from the intercept
because the change of temperature occurs due to thermal
and the slope of the straight line adjusted to the plot of ln(k)
exchange between the reactor and the environment, since all
versus 1/Tr:
limiting reactant has been consumed. For this temperature
decrease, the governing equation is E1
lnk lnk0  16
R Tr
dTr
mCp r UATr Tamb 12 In Eq. (14), the values of the derivative dTr/dt are calculated by
dt
nite difference method, using the following 2nd order central
Using the integrated form of Eq. (12), the slope of a plot of
nite difference or three-point differentiation formulas:
ln(Tr  Tamb) versus t corresponds to UA/(mCp)r, from which the
global heat transfer coefcient can be determined. initial point

Thus, given UA and the known parameters (  DH) and (mCp)r, dTr  3Tr t1 4Tr t2 Tr t3 17
the conversion of acetic anhydride can also be estimated by 
dt t0
2Dt
combining the mass and energy balances and integrating the
resulting equation from t0 to tf assuming that the conversion of
interior points
limiting reactant is equal to 100% at the end of the experiment. 

Therefore, the experimental values of conversion of acetic dTr  Tr tj 1 Tr tj1 18

anhydride are given by the following recursive relationship: dt  2Dt
jt
Rt
mCp r Tr,i 1 Tr,i UA tii 1 Tr Tamb dt
XAi XAi 1  13 last point
DHNA0 
dTr  Tr tn2 4Tr tn1 Tr tn 19

dt t 2Dt
Table 1 n

Room temperature.

Reactor Room temperature Tamb (K)


3. Experimental
1 294.65
2 295.15 The hydrolysis of acetic anhydride was carried out with excess of
3 294.45
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.
Table 4
Volume (mL) Mass (g) Specic heat capacity of materials of reactors and stirrer.

Deionized water 150 Cp (J/g K) Ref.


Acetic anhydride 90
Cylindrical plastic vessel 6.68 Cylindrical plastic vessel 1.79 Ullmann (1985)
Volumetric ask 80.10 Volumetric ask 0.75 Ullmann (1985)
Thermal bottle 23.59 Thermal bottle 0.75 Ullmann (1985)
Magnetic stirrer 4.40 Magnetic stirrer 0.65 Westrum and Grnvold (1969)

Table 3
Constants for specic heat capacity of reactants and products.

A (J/mol K) B (J/mol K2) C (J/mol K3) D (J/mol K4) Ref.

2 4 7
Water 92.053  3.995  10  2.211  10 5.347  10 Yaws (1999)
Acetic anhydride 71.831 8.888  10  1  2.653  10  3 3.350  10  6 Yaws (1999)
Acetic acid  18.944 1.0971  2.892  10  3 2.93  10  6 Yaws (1999)
ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858 3853

reaction, the reaction medium was kept well stirred by a magnetic Table 6
stirrer, and the reactor surrounds were at room temperature. As the Initial conversions.
reactions were carried out on different days, the room temperatures
Reactor XA at t 0 (%)
were different for each type of reactor and are reported in Table 1.
1 6.33
2 4.94
3 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 0.424
2 0.280
3 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 4. Results and discussion
experiments with water (no reaction) for longer periods and no
detectable change in temperature was measured. Thus, the heating In order to determine experimental values of activation energy
effect of stirring was neglected. and frequency factor, the global heat transfer coefcient must be
The rst two systems were used to simulate a non-adiabatic known. This coefcient can be experimentally inferred from
reactor while reactor 3 simulated a non-ideal adiabatic reactor. analysis of temperature variations during the temperature
These systems were chosen to allow the implementation of this decrease where no reaction occurs due to complete conversion
study as an experiment for the undergraduate laboratory. The of the limiting reactant. Taking the integrated form of Eq. (12), UA
temperatures were recorded every 30 s using a digital thermo- can be readily estimated from the slope of a plot of ln(TrTamb)
meter with a sheathed thermocouple. versus t (Fig. 1). The global heat capacity of the reactor, used in
Acetic anhydride having a minimum assay of 95% and Eq. (12), was computed by considering the heat capacity of the
deionized water were used and the volumes of both are shown reactor and the magnetic stirrer plus the average heat capacities of
in Table 2 along with masses of the vessels and the magnetic water and acetic acid produced by complete conversion of acetic
stirrer (magnetiteFe3O4). anhydride. The estimated values of UA are presented in Table 5.
The specic heat capacity of the reactants and products was The correlation coefcients in Fig. 1 ranged from 0.997 to 0.999.
calculated by Comparing the estimated values of heat transfer coefcients
presented in Table 5, it can be seen that the value of UA for the
Cp A BT CT 2 DT 3 20 volumetric ask is slightly lower than that obtained for reactor 1.
This occurs because the heat transfer resistance of reactor 2 is
The constants AD are presented in Table 3. Specic heat greater than that of the polystyrene cup. Besides, the long neck of
capacities of the reactor wall and the magnetic stirrer are the volumetric ask makes heat exchange between the reaction
presented in Table 4 and the heat of reaction for the hydrolysis medium and the environment difcult.
of acetic anhydride was assumed to be (  DH) 58,994.4 J/mol It is important to note that while heat exchange through the
(Conn et al., 1942). double walls of the thermal bottle is minimal, reactor 3 presents a
In Table 1 the mass of the thermal bottle was calculated nonzero value of the heat transfer coefcient, because the thermal
disregarding the external wrapper and considering that the mass bottle was kept open throughout the reaction, allowing for some
of glass ampoule corresponds to 10% of the total mass of the heat exchange though the open top of the bottle.
bottle. The external wrapper was not considered because we Using the estimated values of UA, the experimental prole for
consider that only the internal wall of the ampoule reaches the conversion can be obtained by recursive application of Eq. (13).
same temperature of the liquid inside the reactor. The specic Because all reactions were carried out with an excess of deionized
heat capacities of the volumetric ask and glass ampoule (Table 3) water, conversions were computed from the nal point, where the
are related to the specic heat capacity of an alloy composed 96% conversion of acetic anhydride was assumed to be equal to 100%.
of silicon oxide. 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
Table 7 from zero. Therefore, apparently the actual purity of the acetic
Experimental activation energy and ln(k0) values for the hydrolysis of acetic
anhydride used during the reactions is slightly smaller than the
anhydride.
nominal purity provided by manufacturer. The decrease in purity
Reactor ln(k0) (L/mol s) E (kJ/mol) may be caused by the duration of storage of the reagent or contact
with the environment since acetic anhydride reacts with moisture
1 16.25 68.9 in air.
2 15.28 66.5
3 15.15 66.3
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 Measuring techniques E (kJ/mol) ln(k0) Units of k0


(mol/L)

King and Glasser (1965) n.r. Calorimetry 39.8 9.93 s1


Eldridge and Piret (1950) n.r. Titration 43.2 7.53 L/(mol s)
Takashima et al. (1971) Innite dilution n.r. 33.6
Cleland and Wilhelm (1956) 0.020.06 Titration 44.4 7.80 L/(mol s)
Dyne et al. (1967) 0.17 Calorimetry 49.4
Bisio and Kabel (1985) 0.22 Calorimetry 46.5 12.80 s1
Glasser and Williams (1971) 0.25 Calorimetry 45.3 7.95 L/(mol s)
Takashima et al. (1971) 0.25 n.r. 49.4 18.1 s1
Shatynski and Hanesian (1993) 0.27 Calorimetry 46.9 12.74 s1
Wilsdon and Sidgwick (1913) 0.34 Conductivity 50.6 18.52 s1
Kralj (2007) 0.41 Conductivity 50.1 14.21 s1
Rivett and Sidgwick (1910) 0.54 Conductivity 43.2
Haji and Erkey (2005) 0.66 FTIR 53.6 15.48 s1
Asprey et al. (1996) 1.0 Conductivity 45.7 7.66 L/(mol s)
Wilsdon and Sidgwick (1913) 1.10 Conductivity 56.2 20.45 s1
Marek (1954) 1.34 n.r. 57.8
Marmers (1965) Equimolar concentration n.r 68.7

n.r. not reported.


ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858 3855

and conversion, both k0 and E can be easily computed graphically with additional literature information. It is important to note that,
via Eq. (16). Fig. 2 shows a typical Arrhenius plot according to in Table 8, some frequency factors were obtained under the
Eq. (16), where the slope is equal to (  E/R) and the intercept (at hypothesis of pseudo-rst-order reaction conditions. Therefore,
1/Tr 0) is ln(k0). As can be seen from Fig. 2, the data fall on a for comparison purposes, these parameters should be converted
reasonably straight line with correlation coefcients above 0.99. to second-order constants by dividing the pseudo-rst-order
The kinetic parameters (k0 and E) obtained for each system are constant by the concentration of water, when available.
summarized in Table 7. The literature values for the activation Comparing the kinetic parameters presented in Table 7
energy and frequency factor for the hydrolysis of acetic anhydride obtained in the present work, it can be observed that the values
are listed in Table 8, based on the values reported by Asprey et al. are very similar, regardless of the reactor used. However, when
(1996) and Shatynski and Hanesian (1993) and complemented 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 17.57 73.74
2 16.65 70.14
3 16.19 69.12

100

80

60
E (kJ/mol)

40 literature data

present work
20
present work, considering
only the data for T > 35 C
0
0 2 4 6 8 10
initial concentration of acetic anhydride (mol/L)

Fig. 3. Arrhenius plot for comparing the rate constant for the hydrolysis of acetic Fig. 4. Effect of initial concentration of acetic anhydride on activation energy of
anhydride obtained in this work with those reported in the literature: (a) reactor 1, the hydrolysis reaction. Each point corresponds to one work from the literature,
(b) reactor 2, and (c) reactor 3. 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 kinetics, and both frequency factor and activation energy increase
mixture during the beginning of the experiment, the kinetic with increasing acetic anhydride concentration. Consequently, for
parameters were also computed using the temperatures mea- the anhydride concentration range studied in the literature
sured only after the initial disturbances disappeared (Tr 435 1C). (0.2911.47 mol/L) no unique value of either ln(k0) or E could be
The new kinetic parameters thus estimated are summarized in taken. Besides, Janssen et al. (1957) noted a decrease in the rate
Table 9. constant in the presence of concentrated acetic acid. This fact can
Comparing the values of activation energy and ln(k0) pre- be attributed to the formation of hydrates of acetic acid, as quoted
sented in Tables 7 and 9, it can be seen that the results are very by Plyler and Barr (1935). According to these authors, the
similar, implying that solubility is not mainly responsible for the hydration of some of the water with acetic acid formed probably
variability in reaction rates. In fact, Golding and Dussault (1978) keeps the water from having a part in the reaction and the
noted that data in the literature indicated that in the case of presence of acetic acid also decreases the number of collisions in a
excess of either reactant there is deviation from second order given time between the anhydride and water molecules. On the

Fig. 5. Experimental and predicted values of temperature for: (a) reactor 1, Fig. 6. Experimental and predicted values of conversion for: (a) reactor 1,
(b) reactor 2, and (c) reactor 3. (b) reactor 2, and (c) reactor 3.
ARTICLE IN PRESS
W.H. Hirota et al. / Chemical Engineering Science 65 (2010) 38493858 3857

other hand, Golding and Dussault (1978) also observed that the V total volume of reactants added to reactor, L
effect of acetic acid on reaction rate reported in the literature has XA conversion of component A
been conicting. (  DH) heat of hydrolysis of acetic anhydride, J/mol
Fig. 4 shows the variation of activation energy obtained by r density, g/L
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
Acknowledgements
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 The nancial supports from FAPESPFundac- a~ o de Amparo a
changes of E with temperature. Pesquisa do Estado de Sa~ o Paulo, CNPqConselho Nacional
Finally, in order to test the values obtained for the 3 de Desenvolvimento Cientco e Tecnologico, and CAPES
parameters (U, k0, and E), they were used in the energy and mass Coordenac- a~ o de Aperfeic-oamento de Pessoal de Nvel Superior
balances to simulate the evaluation of temperature and conver- are gratefully appreciated.
sion. Figs. 5 and 6 show a comparison between the experimental
values of the reactor temperature (Fig. 5) and the conversion References
(Fig. 6) with those predicted by the energy and mass balances
using the estimated values of UA and kinetic parameters (k0 and E) Ampelli, C., Di Bella, D., Lister, D.G., Maschio, G., Parisi, J., 2003. The integration of
for each case analyzed. For all cases, a good agreement between an ultravioletvisible spectrometer and a reaction calorimeter. Journal of
Thermal Analysis and Calorimetry 72, 875883.
experimental and predicted values was found despite values of Ampelli, C., Di Bella, D., Lister, D.G., Maschio, G., 2005. Fitting isoperibolic
activation energy and frequency factor obtained being always calorimeter data for reactions with pseudo-rst order chemical kinetics.
higher than those reported in literature. 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.
5. Conclusions 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
This work shows that both kinetic parameters and heat spectroscopy. Analytical Chemistry 70 (2), 332339.
Bisio, A., Kabel, R.L., 1985. Scaleup of Chemical Process. Wiley, New York (pp. 136138).
transfer coefcient of reactor contents can be easily determined Cleland, F.A., Wilhelm, R.H., 1956. Diffusion and reaction in viscous-ow tubular
for hydrolysis of acetic anhydride using temperature-versus-time reactor. AIChE Journal 2, 489.
data under non-adiabatic, isoperibolic conditions. The main 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
difference of this work in relation to other previously published
American Chemical Society 64 (8), 17471752.
works in the literature is that the kinetic data were obtained Dyne, S.R., Glasser, D., King, R.P., 1967. Automatically controlled adiabatic reactor
under non-adiabatic conditions, where the heat transfer coef- for reaction rate studies. The Review of Scientic Instruments 38 (2), 209214.
cient must be known in order to determine these parameters. The Eldridge, J.W., Piret, E.L., 1950. Continuous-ow stirred-tank reactor system I.
Design equations for homogeneous liquid phase reactions. Experimental data.
results shows that the estimates obtained for activation energy Chemical Engineering Progress 46, 290.
and frequency factor of the hydrolysis of acetic anhydride are Glasser, D., Williams, D.F., 1971. The study of liquid-phase kinetics using
apparently affected by excess of either reactant and, therefore, no temperature as a measured variable. Industrial and Engineering Chemistry
Fundamentals 10 (3), 516519.
unique value of either ln(k0) or E could be taken. Despite the Gold, V., 1948. The hydrolysis of acetic anhydride. Transactions of the Faraday
observed differences between estimates of k0 and E and those Society 44, 506518.
reported in the literature, experimental and predicted values of 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
temperature and conversion show good agreement. Chemical Engineering 56, 564569.
As a nal comment, the simplicity of the experimental setup Haji, S., Erkey, C., 2005. Kinetics of hydrolysis of acetic anhydride by in-situ FTIR
and the calculations used in this investigation allows one to use spectroscopy. An experiment for the undergraduate laboratory. Chemical
Engineering Education 39 (1), 5661.
them as an experiment for undergraduate laboratories to
Janssen, H.J., Haydel, C.H., Greathouse, L.H., 1957. Hydrolysis of acetic anhydride in
illustrate the application of the non-adiabatic calorimetric concentrated acetic acid without catalysis. Industrial and Engineering
method to infer state variables. 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
Notation conductivity. Journal of Industrial and Engineering Chemistry 13 (4), 631636.
Marek, J., 1954. Collection of Czechoslovak Chemical Communications 19, 621.
CA concentration of acetic anhydride, mol/L Marmers, H., 1965. Ph.D. Dissertation, University of Birmingham, Birmingham,
CW concentration of water, mol/L England.
Oakenfull, D.G., 1974. The kinetics of the hydrolysis of acetic anhydride and the
Cp specic heat capacity, J/mol K or J/g K reaction of 4-nitrophenyl acetate with imidazole in aqueousorganic mixed
E activation energy, J/mol solvents. Australian Journal of Chemistry 27 (7), 14231431.
k specic reaction rate constant, L/mol s Orton, K.J.P., Jones, M., 1912. Hydrolysis of acetic anhydride. Journal of the
Chemical Society 101, 17081720.
k0 pre-exponential factor or frequency factor, L/mol s Plyler, E.K., Barr, E.S., 1935. The reaction rate of acetic anhydride and water. Journal
(mCp)r total heat capacity of reactor and contents, J/K of Chemical Physics 3 (11), 679682.
MW molecular weight, g/mol Poc- o, J.G.R., Danese, M., Giudici, R., 2010. Investigation of cationic polymerization
of beta-pinene using calorimetric measurements. Macromolecular Reaction
NA0 moles of acetic anhydride initially fed to reactor, mol
Engineering 4 (2), 145154.
NW0 moles of water initially fed to reactor, mol Puxty, G., Maeder, M., Rhinehart, R.R., Alam, S., Moore, S., Gemperline, P.J., 2005.
R gas constant, J/mol K Modeling batch reactions with in situ spectroscopy measurements and
calorimetry. Journal of Chemometrics 19, 329340.
( rA) rate of disappearance of component A, mol/L s
Regenass, W., 1985. Calorimetric monitoring of industrial chemical process.
t0 initial instant, s Thermochimica Acta 95, 351368.
tf nal instant, s Rivett, A.C.D., Sidgwick, N.V., 1910. The rate of hydration of acetic anhydride.
Tamb environment temperature, K Journal of the Chemical Society 97, 732741.
Shatynski, J.J., Hanesian, D., 1993. Adiabatic kinetic studies of the cytidine/acetic
Tr reactor temperature, K anhydride reaction by utilizing temperature versus time data. Industrial and
UA heat transfer coefcient, J/K s 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 Wilsdon, B.H., Sidgwick, N.V., 1913. The rate of hydration of acid anhydrides: acetic,
Physical Chemistry 59 (5), 385388. propionic, butyric, and benzoic. Journal of the Chemical Society 103, 19591973.
Takashima, I., Nishida, A., Ogita, K., Uchida, N., 1971. Development of liquid phase Wojciechowski, B.W., 1997. The temperature scanning reactor I: reactor types and
reaction analyser 3. Computer aided reaction - analysis on twin calorimetry. modes of operation. Catalysis Today 36, 167190.
Kogyo Kagaku Zasshi 74, 12931297. Yaws, C.L., 1999. Chemical Properties Handbook: Physical, Thermodynamic,
Ullmann, F., 1985. Ullmanns encyclopedia of industrial chemistry, 5th ed Wiley- Environmental, Transport, Safety, and Health Related Properties for Organic
VCH, Weinheim. and Inorganic Chemicals. McGraw Hill, New York.

Visentin, F., Gianoli, S.I., Zogg, A., Kut, O.M., Hungerbuhler, K., 2004. A pressure-
Zogg, A., Fischer, U., Hungerbuhler, K., 2003. A new small-scale reaction
resistant small-scale reaction calorimeter that combines the principles of calorimeter that combines the principles of power compensation and heat
power compensation and heat balance. Organic Process Research and balance. Industrial and Engineering Chemistry Research 42 (4), 767776.
Development 8 (5), 725737.
Zogg, A., Fischer, U., Hungerbuhler, K., 2004. A new approach for a combined
Westrum, E.F., Grnvold, F., 1969. Magnetite (Fe3O4) heat capacity and thermo- evaluation of calorimetric and online infrared data to identify kinetic and
dynamic properties from 5 to 350 K, low-temperature transition. The Journal thermodynamic parameters of a chemical reaction. Chemometrics and
of Chemical Thermodynamics 1 (6), 543557. Intelligent Laboratory Systems 71 (2), 165176.

You might also like