Вы находитесь на странице: 1из 8

Ind. Eng. Chem. Res.

1994,33, 277-284

277

Kinetics, Mass Transfer, and Palladium Catalyst Deactivation in the


Hydrogenation Step of the Hydrogen Peroxide Synthesis via
Anthraquinone
E. Santacesaria,J M.D i SerioJ R. Velotti? a n d U. Leonet
Dipartimento di Chimica, Uniuersith di Napoli, Via Mezzocannone 4, 80134 Napoli, Italy, and
Ausimont SPA-CERZOS, Bussi (Pescara),Italy

Hydrogenation of 2-ethyl-5,6,7,8-tetrahydroanthraquinone
is a key step in the industrial production
of hydrogen peroxide via anthraquinone with the process named all-TETRA. This reaction on
palladium-supported catalysts is very fast; consequently, the chemical regime can hardly be achieved
and kinetics is always masked by mass-transfer limitations. Nevertheless, it is possible to demonstrate
that the reaction occurs with a zero- and first-order kinetics with respect to hydrogen and to the
organic reagent, respectively. These reaction orders can be explained on the basis of reasonable
reaction mechanisms described and discussed in this paper. Kinetics has been studied by performing
runs in two different laboratory reactors: a semibatch and a continuous stirred tank reactor. In
particular, the continuous reactor has been used for studying catalyst deactivation. Two types of
catalyst poisoning have been recognized, a reversible one, related to the presence of water, and a
permanent, not yet explained one. A kinetic expression is given also for deactivation. The kinetic
parameters obtained from the experimental runs have been verified by simulating the behavior of
an industrial reactor also considering catalyst deactivation.
Introduction
In the process named all-TETRA (see Powell, 1968;Kirk
and Othmer, 1981; Ulmanns, 1989) for the industrial
production of hydrogen peroxide, the hydrogenation of
2-ethyl-5,6,7,8-tetrahydroanthraquinone
(THEAQ) is a
fundamental step. In fact, reactions occurring in the
process are the following:

II

?H

OH

(THEAQ)

(THEAOH,)

OH

0
(THEAQ)

Hydrogen peroxide is then extracted with a slightly acidic


solution. Although the working solution normally contains
30% ethylanthraquinone (EAQ) and 70% THEAQ (Powell, 1968;Kirkand Othmer, 1981;Ulmanns, 1989),THEAQ
only participates in the production cycle because hydrogenation is usually kept at about 70% of conversion and
the following equilibrium is operative:
EAQH, + THEAQ F? EAQ + THEAQH,
(3)
The equilibrium is completely shifted to the right, as
reported from Berglin and Shoon (1983).Hydrogenation
of THEAQ occurs both directly and via EAQ. However,
the kinetic behavior of the two mentioned anthraquinones
in the hydrogenation is quite similar. Therefore, their
mixtures can be considered as a single pure component.

* To whom correspondence should be addressed.

+ Universita di Napoli.

* Ausimont SPA-CERIOS.

Hydrogenation is normally performed in slurry reactors,


in the presence of a palladium-supported catalyst. Very
few papers have been devoted to the kinetics of this
reaction. Kirdin et al. (1970)and Berglin and ShMn (1981),
for example, studied the reaction in the presence of RaneyNi catalyst. Santacesariaet al. (1988)studied the kinetics
and mass transfer in the presence of a palladium-supported
catalyst and concluded that palladium is extremely active
and the reaction is always controlled by mass transfer.
Therefore, kinetic data were interpreted in that work by
neglecting the chemical reaction rate contribution. This
assumption is a rough approximationwhich can reasonably
be accepted only for runs performed with fresh catalysts.
However, it was observed that palladium catalysts are
susceptible to deactivation by poisoning. Therefore, the
importance of the chemical reaction rate contribution can
be estimated by observing the effect of the catalyst
poisoning. The present work must therefore be considered
an improvement of the one published from Santacesaria
et al. (1988).Reaction orders and kinetic parameters will
be determined, the kinetic law found will be discussed,
and the related reaction mechanism will be suggested.
Kinetic parameters have been determined from runs
performed in laboratory slurry semibatch and continuous
reactors.
Mass-transfer parameters have newly been estimated,
and the obtained values are in good agreement with those
reported in the previous work (Santacesaria et al., 1988).
Catalyst deactivation kinetics has been studied in a
laboratory continuous reactor. Two deactivation mechanisms have been recognized, a reversible and fast one
due to the presence of water and a permanent and slow
one probably due to the formation of bulky molecules on
the palladium surface as a consequence of anthraquinone
molecules condensation. The kinetic model developed in
this work, also considering catalyst deactivation, has
successfully been tested in an industrial plant.
Experimental Section
Apparatus, Methods, and Reagents. Kinetic runs
were performed in a slurry semibatch reactor. The scheme

0888-5885/94/2633-0277$04.50/0 0 1994 American Chemical Society

278 Ind. Eng. Chem. Res., Vol. 33, No. 2, 1994

w
Automatic

Figure 1. Scheme of the slurry semibatch reactor. A = hydrogen


feed, tube is connected to a mass flowmeter and to the hydrogen
bottle; B = exit for liquid sample withdrawing: C = magnetic drive
stirrer; D = baffles; E = inlet of thermostatic fluid; F = outlet of
thermostatic fluid; G = thermal insulation; TC = thermocouple.

Figure 2. Scheme of the continuous reactor. A = hydrogen feed; B


= feed of solution; C = magnetic drive stirrer; D = baffles; E = inlet
of thermostatic fluid; F = outlet of thermostaic fluid; G = wire cloth;
H = solution outlet; LC = liquid level automatic control: S = sensing
element.
cm3 of HZ adsorbed

of the reactor is reported in Figure 1. A solution containing


the mixture of EAQ (30%) and THEAQ (70%) commonly
used in industry has been employed in all the experiments.
The concentrationsof the reactants have been determined
by gas chromatographic analysis performed on a Cp-Sil19capillary column of 25-m length and 0.25-mm diameter.
As the resulting kinetic behavior of the two mentioned
substances was quite similar, we will treat the mixture as
a unique active component called "anthraquinone". The
stainless steel reactor was jacketed and equipped with a
magnetically driven stirrer consisting of a turbine connected, as in Figure 1, to a hollowed cylinder and able to
develop a great interfacial area. Teflon baffles were put
inside the reactor. The internal size of the reactor was
exactly the same as that of the glass reactor described in
a previous paper by Santacesaria et al. (1988). Therefore,
the reactor fluid dynamics was well-known. Hydrogen
consumption was directly measured with a mass flowmeter
connected to a computer. Thus, we evaluated the instantaneous hydrogen flow rates, that is, the reaction rates
and the total volume of reacted hydrogen at different times.
Kinetic runs were performed by changing catalyst concentration, rotating speed of the stirrer, anthraquinone
concentrations, and temperature.
Other kinetic runs were performed in a laboratory
continuous reactor whose scheme is reported in Figure 2.
This reactor has the same fluid dynamic characteristics
of the already described semibatch reactor of Figure 1,
being identical in the design of the stirrer, the baffles, and
the internal size. A closely woven wire net keeps the
catalyst confined inside the reactor. Again, reaction rates
were determined by measuring hydrogen consumption
rates with a mass flowmeter connected with a computer.
The kinetic and mass-transfer parameters determined in
this work reproduce well the data obtained with both the
semibatch and continuous reactors. However, the continuous reactor has mainly been used to study catalyst
deactivation in long time runs.
The catalyst employed and solvent characteristics are
the same as reported by Santacesaria et al. (1988).

3500 1

I-

2ooo
1500-

.:::I

500 1000 1500 2000 2500 3000 3500 4000 4500


time

(s)

Figure 3. Kinetic run performed at 70 O C , speed = lo00 rpm, m =


0.005 g/cm3, [THEAQ] = 4.7 X le mol/cm3,p
H
,
'
1
atm, V ~ = 2 0 0
cm3.
Table 1. Main Properties of the Catalyst Used
average diameter
specific surface area
metallic surface area
specific weight
bulk density
porosity

0.013 cm
178 m2/g
149 m2/g
2.25 g/cm3
0.77 g/cm3
0.80 cm3/g

However, some catalyst properties are summarized in


Table 1. All the reagents were provided by Ausimont SPA,
while the catalyst was supplied by Montecatini Tecnologie
SPA. The stainless steel reactors were built by INOX
Impianti Co.
Results and Discussion
Kinetic Runs in the Semibatch Reactor and Their
Interpretation. Figure 3reports the example of a hydrogenation run performed at 70 "C and 1000 rpm with
a 0.0055 g/cm3 catalyst concentration. In this run the
hydrogen consumption trend during time is nearly linear

Ind. Eng. Chem. Res., Vol. 33, No.2, 1994 279


cm3 of H, adsorbed

%JHr

(5)

I
500

40

180

120

60

time

3 20

so0

240

700 rmn/

(5)

Figure 4. Kinetic N ~ Sperformed at different catalyst concentrations. Catalyst hold-up is reported on the plot in gIcm3. Temperature=% 'C, speed = loo0 rpm, [THEAQ1=0.88 x lo-' mollcms,
PH,=1 atm, V~=300
ems. Symbols correspond to experimentaldata;
lines are obtained by calculations.

10

0
0

100

200

300

4w

500

l / m (cm Yg)
Figure 6. Trend of pHJ" against l/m at different stirring ratea.
Temperature=%C,[THEAQI =O.&?X l ( r m o l l c m s , p ~ ~ l a t m .

vL=300 cm3.

300

*.-'

200

d
d

,.*."

60

120

loo"d",,'./
o,<p'--po

,
.
"""

180

240

_1
JOO

JBO

120

480

$40

sponding to a constant reaction rate. In agreement with


the suggestions of Chaudari and Ramachandran (1980),
this behavior can be explained with a zero-order kinetics
with respect to the gaseous reagent.
In fact, when a first-order kinetics is valid we can write
the following steps:

NGL= kLa,[(pH /H)


- C,] = gas-liquid mass transfer
(4)

time (SI

Figure 5. Kinetic runs performed at different catalyst concentrations. Catalyst hold-upisreportedon tbeploting/cms.Temperature
= 70 "C, speed = loo0 rpm, [THEAQl = 4.7 X l(r moVcms,PH, =
1atm, VL=ZOOems.Symbols correspondto experimentaldata; lines
are obtained by dculationa.

for about 60% of the substrate conversion and then


gradually decreases. In the last part of the run, the linear
behavior disappears and reaction rates become sensitive
to the organic reagent concentration. Therefore, we
a(rsumed a firsborder kinetie with respect to anthraquinone as a first approximation: as will be seen, all the kinetic
runs are well fitted with this assumption.
Many kinetic runs have been performed at two different
temperatures of 50 and 70 "C by changing catalyst
concentration and the rotating speed of the stirrer. In
Figures 4 and 5, for example, runs performed respectively
at 50 and 70 "C and 1000 rpm, in the presence of different
catalyst concentrations, are reported. Anthraquinone
concentration has been taken 4.71 X lo-' mol/cm3 at 70
' C and 0.88 X lo-' at 50 "C. Data of Figures 4 and 5 have
been rearranged in plots showing the trends of pH,/Hr
against l / m (see the Nomenclature section and relation
8), as reported in Figures 6 and 7. Figure 6 shows the
obtained values of pHJHr against llm for runs performed
at5OoCandatthreedifferentrotatingspeedsofthestirrer,
Le., 700,1000, and 2000 rpm. Figure 7 reproduces runs
similarly performed at 70 O C .
As can be seen from Figs 6 and 7, for higher catalyst
concentrations a plateau is obtained in all cases, corre-

NM

= ksas(C, - C,) = liquid-solid mass transfer

r = q p k , C s = reaction rate

(5)
(6)

By assuming pseudo-steady-state conditions, the three


relations can be equalized, and CI and Cs eliminated.
Remembering that as = 6m/p,d,, we can write

r=Tqgqi
p,/H

sP1 m

(7)

k ~ a ~
Rearranging the expression, we have

that is, a linear correlation between pHdHr and l/m.


In addition, if ppdd6ks >> l / q ski, the last term could
be neglected as has been done in a previous work by
Santacesaria et al. (1988). However, in Figure 7, we
compare the kinetic behavior in a batch reactor of a fresh
catalyst with that of an aged catalyst having an average
life time of 100 days in an industrial plant. As the mean
size of the particle remained unchanged, the difference in
the slopes observed must be attributed to the difference
in the kinetic constant as a consequence of poisoning.
Therefore, chemical contribution cannot be neglected in
the case of an aged catalyst.

280 Ind. Eng. Chem.

Res., Vol. 33, No.2, 1994

p H p (8)
I

1-6

Figure 9. Concentrationprofiles for a zemorder reaction. (a) Case


in which q = 1; (b) case in which q <l. (See Chaudari and
Ramachandran (1980).)

01
0

100

2QO

300

400

I
500

l / m (cm Yg)

Figure 7. Trend ofpHJHragainstl/m for fresh (m) and aged catslyst


(A).Temperature =IO 'C, speed = 1oW rpm. [THEAQI = 4.7 X 1W
mol/cms, PH, = 1 atm, VL = 200 cma.
PAI H r (s)

rates increase until the limiting rate of the external mass


transfer is reached. Any further catalyst addition only
has the effect of reducing the working catalyst portion
more and more.
A zero-order kinetics with respect to hydrogen can easily
be explained by considering the activity of palladium in
the hydrogen adsorption. First of all, we will have a
dissociative adsorption of the type
H, + 20 a 2Ha
(9)
with equilibrium strongly shifted to the right. First order
with respect to anthraquinone can be explained by
assuming for this reaction a Rideal mechanism of the type

THEAQ 2Ha THEAQH, + 20


The hyrogenation rate should, therefore, be

(10)

If (~H,CH,)'/~
>> 1, then
1lm (cm3 /gr)

FigureO. Typicaltrends ofpdHragainst l l m for differentreaction


orders in which a gaseous A reagent is involved. (See Chaudari and
Ramachandran (1980).)

When the reaction order with respect to the gaseous


reagent is different from 1, we do not have a linear plot
for PHJfh' against l / m and the behaviors are different
according to the reaction order, as can be seen in Figure
8 (Ramachandran and Chaudari, 1983).
By comparing Figure 8 with Figures 6 and 7, we must
conclude that anthraquinonehydrogenation is a zero-order
reaction with respect to hydrogen. The plateau observed
in the plots of Figures 6 and 7, typical for a zero-order
reaction, as seen from Figure 8, can be explained by
considering that, with the reaction rate independent of
the gaseous reagent concentration, this concentration can
easily be zeroed inside the catalyst particles before reaching
the particle center, especially in the case of sparingly
soluble gases. We can have two possible situations: (i)
the gaseous reagent concentration is never zeroed inside
the particle and the catalyst efficiencyis 1;(ii)the gaseous
reagent concentration is zeroed before the center of the
particle is reached and the catalyst effectivenessis always
lessthan 1. These twosituationsarerepresentedin Figure
9. When the second possibility is operative, only a portion
ofthecatalyst works. By additionofthecatalyst, reaction

r = mk[THEAQI
(12)
In order to interpret kinetic runs, we must consider both
mass-transfer and reaction steps, that is,

NGL= kLaL[@%//H)
- Cll

(13)

NH = ksas(Cl - C,)

(14)

r = smk[THEAQl
(15)
Ramachandranand Chaudari (1983) described the way to
calculate the overall effectiveness factor for a zero-order
reaction. In our case, as the THEAQ concentration is
always much higher than the hydrogen one, it is reasonable
to consider the THEAQ concentration constant for a very
small time and to assume, therefore,

k[THEAQI = ko -,
pseudo-zero-order constant (16)
Thus, the model developed by Chaudary and Ramachandran can directly be applied to o w reaction system, too.
Moreover, in o w experiments the calculated effectiveness factor was always less than 1. Calculations have been
made with the relation (see Chaudari and Ramacbandran,
1980)

n = VDW- (8/6)1[1- 3(1 - d 2 / 3+ 2(1 -?)I)

(17)
where w is the overall effectiveness factor for a system
completely dominated by external diffusion. In this case;

Ind. Eng. Chem. Res., Vol. 33, No. 2, 1994 281


Table 2. Kinetic and Mass-TransferParameters

the reaction rate is

~~~

700
lo00
2000
1000

0.10
0.07
0.11
0.28
0.08
0.27
50
1.4
0.78
0.13
0.75
70
3.0
0.32
0.12
Parameters obtained by regression on experimental data. * Parameters obtained by physical measurements (&) or by calculations
(ks).
50
50

and

It is interesting to observe that for a very fast reaction r]


is negligible with respect to 1;consequently according to
. could be a small value considering
relation 17 r] = 7 ~ VD
that k, in relation 19 could be very high with respect to
the numerator.
The first calculation step is the evaluation of the 4
modulus (see Chaudari and Ramachandran, 1980):

1.4
1.4

Ks (cmls)

0'25
0.2

q
-

70C

0.15.

50'C

0.1

Then we calculate ?JDwith relation 19 and r] by solvingthe


equation

- (4'/6)1[1-

3(1- q ) 2 / 3+ 2(1 - 7)1) = 0


(21)
with the secant method (see Norris, 1981).
At last, the reaction rate can be calculated as

F(r])= r] - qD{[1

r = qmk,
(22)
Batch runs can be simulated by integrating the following
differential equations:

d V&
-- VMVLk,as(C,- Cs) = r"
dt

30C

0.05

goo

1000

1500

2000

2500

rpm

Figure 10. Values of ks calculated at different temperatures and


rotating speed rates of the stirrer.

described in detail by Santacesaria et al. (1988). According


to the second method, PL can be determined from
experimental plots such as those reported in Figures 6
and 7 from the value of pH,/Hr at the intercept. As a
matter of fact, the value of the intercept corresponds
approximately to

(24)

Vi,, V: ,and VEt are the volumes of hydrogen adsorbed


by gas-liquid mass transfer, liquid-solid mass transfer,
and reaction, respectively. 7 is calculated as already
described, and k , is adjusted at any step of integration
remembering that

k , = k[THEAQ] = k[THEAQl0(1- X)
(28)
where [THEAQI O is the initial anthraquinone concentration and X is its conversion.
In order to make calculations, we need to know the
following parameters: PL = kLuL and Bs = ksas and their
dependence on temperature as well as stirring rate; De,
hydrogen solubility H , kinetic constant k, and their
dependence on temperature.
BL values have been determined in two distinct ways.
The first way is based on the measurement of the evolution
of the interface area with the stirring speed using the sulfite
method as suggested from Link and Vacek (1981) and as

Values obtained according to the two ways described are


compared in Table 2. As can be seen, the agreement is
satisfactory. As /3s = ksas, with as= 6mlqpdp,this value
can be determined from the geometrical characteristics of
the catalyst particles summarized in Table 1. The metallic
surface area has been determined by the CO chemisorption
method. ks has been determined with the semiempirical
relation suggested by Sano et al. (1974). The values
obtained for ks at different temperatures and rotating
speed rates of the turbine are reported in the plot of Figure
10. The hydrogen effective diffusion coefficient De has
been estimated on the basis of the characteristics of the
catalyst reported in Table 1 and on the basis of the
molecular diffusion coefficient. On the basis of the values
given by Berglin and ShMn (1981) for the diffusion of
hydrogen in an equimolecular solution of xylene and
2-octanol, for the hydrogen diffusion in the working
solution, we estimated a relation of the type

DH,= 1.05 X

exp(-1520/T) (cm2/s)

(30)

with the aid of the Wilke-Chang (1955) relationship,


considering the differences in the employed working
solutions.
Hydrogen solubility, density, and viscosity of the
working solution as a function of temperature are the same
as those reported by Santacesaria et al. (1988). A t last,
the kinetic parameters have been determined by fitting
all the available kinetic runs; also the runs reported from

282 Ind. Eng. Chem. Res., Vol. 33, No. 2, 1994


cm 3 d H, adlorbed

5~
0
0

60

120 180 240 300 360 420 480 540 600 660 720 780

time (day)

tcme (P)

Figure 11. Kinetic run performed on catalyst having particles with


different mean diameter. Temperature = 70 O C , m = 0.01 p/em3,
speed = loo0 rpm, [THEAQI = 4.7 X lW mol/cm3,PH, = 1 atm,
V ~ = 2 0 0em3. Symbols correspond to experimental data; linea are
obtained by calculations.

Figure 13. Rates of hydrogen adsorption in a continuow reactor


againat time for two different catalyst hold-up (g/cmJ). Water
concentration =3 g/L, temperature = 70 "C, speed = loo0 rpm,
[THEAQ] = 4.7 X 1W mol/cm3,PH,= 3 atm, VL = 300 ems, QL =
1.16 em3/s. Symbols correspond to experimental data; lines are
obtained by calculations.

consumption with a flowmeter and the anthraquinone


conversion at the reactor outlet.
The following mass balance equation has been used for
interpreting continuous kinetic runs:

rl

0.01

0.02

0.03

0.04

0.05

m (dem5

Fiure 12. Evolution of the effectiveness factor against catalyst


hold-up for different temperatures and stirring speed ratea.

Santacesaria et al. (1988) have been reinterpreted. The


parameters obtained are reported in Table 2.
Examples of the fittings obtained by calculations are
reported in Figures 4 and 5. Runs have also been
performed on the same catalyst having particles with a
mean diameter of 0.2 and 0.44 mm instead of 0.13 mm.
The fitting obtained with the parameters of Table 2 can
be appreciated in Figure 11.
In Figure 12, the evolution of the effectiveness factor
with the catalyst hold-up is reported for two different
temperatures (50 and 70 "C). In Figure 12 is also given
the effectivenessfactor at 50 "C for three different stirring
speed rates (700,1000, and 2000 rpm). As can be seen, the
effectiveness factor is always very low in agreement with
the high catalytic activity of the palladium catalyst. As
a consequence, only a thin shell of the catalyst particles
functions normally.
Kinetic Runs in a Continuous Reactor. As mentioned, the continuous reactor of Figure 2 has the same
internal size, baffles, and stirrer as the semibatch reactor
of Figure 1. As a consequence, the preliminary runs
performed in the reactor of Figure 2 in semibatch
conditionsgaveexactlythesameresultsasthoseperformed
in the reactor of Figure 1.
Continuous kinetic runs have been performed with
different liquid feed rates by controlling the hydrogen

Q,([THEAQli- [THEAQI) = smk[THEAQIVL (31)


The kinetic and mass-transfer parameters used in the
calculationsarethesamereportedinTable2. Thecatalyst
effectivenessfactor has been determined with the already
described procedure. A comparison for fresh catalyst
betweenexperimentalandcalculated values together with
the adopted operative conditions can be appreciated in
Figure 13 (see values at zero time). As can be seen, the
agreement is satisfactory in this case, too.
The wire net used for retaining the catalyst inside the
reactor was very efficient considering that after long time
the weight of the catalyst remained unchanged. Longtime kinetic runs have been performed in the continuous
reactor in order to study the catalyst deactivation through
the observation of the decrease with time in hydrogen
consumption.
We recognized that water is responsible for a reversible
poisoning of the catalyst. The working solution isnormidly
saturated by water as a consequence of the extraction of
hydrogen peroxide and contains about 3 g/L water, at 70
"C. By feeding this solution in the continuous reactor, we
obtained an activity plot as reported in Figure 13. In this
plot, we observe an initial fast deactivation and a successive
slow activity stabilizationat 30 % lower level than the initial
one.
Ifwaterisfedinagreateramount,byfeedingforexample
a water suspension, deactivation occurs much faster and
residual activity much lower as can be seen in Figure 14,
where a solution containing 3.7 g/L water is first admitted
to the reactor and then a suspension containing more than
5 g/L water is fed for a prefixed time. When the feeding
of the saturated solution was restored, activity slowly
increased at the expected level. This behavior clearly
shows the reversible nature of catalyst poisoning by water.
On the basis of these observations, an opportune kinetic
model for deactivation can be developed to simulate runs
such as that reported in Figure 13. The same model could
also be used to simulate the performance of an industrial
plant. However, in order to describe the industrial plant
behavior, another irreversible poisoning effect must be
considered.

glL

NL/h of H2adsorbed

30

25

20

I:

15
10

--

40m

20
0

8 10 12 14 16 10 20 22 24 26 28 30
time (day)

Figure 15. Simulation of an industrial reactor. Dashed line


corresponds to the simulation of anthraquinone conversions with a
model neglecting deactivation. Continuous line interpolating experimental points is the simulation achieved by considering the
occurrence of catalyst deactivation. Heavy continuous line corresponds to the percent of increasing catalyst hold up in the plant.
Table 3. Kinetic and Equilibrium Parameters for
Deactivation

ki (cm3/(molday))
6.5 x 103

If 6 is the fraction of free sites,

4)
- free active sites at time t

6 ( t )= UO

(33)

initial total sites

and Ooisthe fraction of available sites, i.e., not permanently


poisoned,
site at time t
e,,(t)= -- total
uo
initial total sites
6* is the fraction of sites occupied by water,

e*=--u*(t) - sites occupied by water =


UO

(34)

(35.

initial total sites

We can put the kinetic constant of the hydrogenation


reaction proportional to 6, that is,

k = k,6

(36)
Initially 6 = 1, then the evolution with time of 6
corresponding to the evolution of k can be interpreted for
the reversible poisoning with the relation

derived from the proposed mechanism reported in relation


32. Irreversible deactivation must be considered together,
and we propose the relation

Initially also 6, = 1. By integrating relations 37 and 38


together, we can simulate runs as the one reported in Figure
13 where the curve is the best fitting obtained with the
parameters given in Table 3. The parameter k: has been
estimated from data of the industrial reactor. In Figure
15, an example is reported of the industrial reactor
simulation. Mass-.transfer parameters were determined

kf (day')
2.75 X 1 t 2

Kd (atm mol/cm3)
3.9 x 104

in this case with semiempirical correlations existing in the


literature (see Akita and Yoshida (1973) and Sano et al.
(1974)). The symbols represent experimental conversions
observed a t different times in corresponding percentage of increasing catalyst hold up. The dotted line is the
simulation obtained by neglecting the deactivation effect.
The mechanism of the permanent deactivation has not
yet been explained but is probably due to the condensation
of two or more anthraquinone molecules on the palladium
surface, although it cannot be excluded that water could
also be responsible for the irreversible poisoning.
Finally, it is interesting to observe that probably a
uniform distribution of palladium inside the catalyst
particles is an advantage because internal palladium works
as catalyst reservoir when the external palladium initially
involved in the reaction is definitively poisoned.

Conclusions
Kinetics and mass transfer in the hydrogenation at the
oxygen of anthraquinones commonly used in industry, such
as THEAQ, EAQ, and their mixtures, have been studied
in the presence of palladium catalyst.
We observed that palladium is a very active catalyst of
this reaction; therefore, mass-transfer limitations strongly
affect reaction rates. In particular, internal diffusion
limitations are often operative, despite the very small size
of catalyst particles used in the experiments (see Figure
12). In order to separate all the possible contributions to
the reaction rates, kinetic law must be identified and
kinetic and mass-transfer parameters must be estimated.
For this purpose, we have shown that the hydrogenation
of anthraquinones on palladium is a zero-order reaction
with respect to hydrogen and first order with respect to
the anthraquinone. The kinetic law has been derived from
both direct and indirect observations. The kinetic constants, reported in Table 2, have been determined by
comparing the activities obtained from different kinds of
reactor, semibatch or continuous, and have been verified
by simulating a continuous industrial plant.

284 Ind. Eng. Chem. Res., Vol. 33, No. 2, 1994

We have also shown that palladium catalysts are


subjected, in this reaction, to both reversible and irreversible poisoning. Water is responsible for the first
poisoning effect, and a quantitative interpretation of the
phenomenon is given.
The irreversible poisoning is too slow to be studied in
a laboratory reactor; therefore, we have interpreted data
collected from a continuous industrial plant during more
than 30 days (see Figure 15). In this plant catalyst holdup has been increased by about 40% in the first 6 days of
operation. Activities of this plant have been successfully
simulated by using all the kinetic and mass-transfer
parameters reported in this paper.
We can conclude, therefore, that our kinetic approach
seems to be reliable, and we hope that our work will be
useful both to optimize the design and management of the
hydrogen peroxide industrial plants and to understand
better the mechanism of hydrogenations on palladium
catalyst to improve the catalyst performance by reducing,
for example, the effect of poisoning.
Acknowledgment
Ausimont SpA is acknowledged for the financial support
of the present work.
Nomenclature
CIL =
US

specific gas-liquid interface area (cm2/cm3)

= specific external area of particles (cmVcrn3)

b~~ = hydrogen adsorption equilibrium constant (cmYmo1)


C H =~ hydrogen concentration in the liquid phase (mol/cm3)
Cl = concentration of gas in the liquid phase (mol/cm3)
CS = concentration of gas at the catalyst surface (mol/cm3)
d , = average diameter of the catalyst particles (cm)
D H =~ diffusion coefficient of Hz in liquid phase (cm2/s)
De = effective diffusion coefficient (cm2/s)
H = Henrys law constant for Hz (atm cm3/mol)
[THEAQ] = total concentration of anthraquinones (mol/
cm3)
[THEAQ], = feed concentration of anthraquinones in the
continous reactor (mol/cm3)
k = rate constant of hydrogenation (cm3/sg)
ko = rate constant of zeropseudo order, ko=k[THEAQ] (mol/
(g 8))
121 = rate constant of first order (cm3/s g)
k, = rate constant of hydrogenation of fresh catalyst (cm3/(s
g))
k i = rate constant of reversible catalyst poisoning (cm3/(mol
day))
kil = rate constant of irreversible catalyst poisoning (l/day)
Ked = equilibrium constant of reversible catalyst poisoning
(atm mol/cm3)
kL = gas-liquid mass-transfer coefficient (cm/s)
k s = liquid-solid mass-transfer coefficient (cm/s)
m = catalyst hold-up (g/cm3)

NG-L= gas-liquid mass- transfer rate (mol/(cm3 8))


N m = liquid-solid mass-transfer rate (mol/(cm38))
r = hydrogenation rate (mol/(cm3e ) )
PI = hydrogen absorption rate at the gas-liquid interface (cm3/
9)

rII = hydrogen absorption rate at the liquid-solid interface

(cm3/s)

rIII = hydrogen consumption as a consequence of the reaction

(cm3/s)

p~~ = hydrogen pressure (atm)


QL = solution feed rates (cm3/s)
R, = mean radius of catalyst particles (cm)
VL = volume of slurry (cm3)
VM = hydrogen molar volume (cm3/mol)
Greek L e t t e r s

= overall gas-liquid mass-transfer coefficient, BL = kLaL


(Us)
PS = ksas = liquid-solid mass-transfer coefficient
7 = overall effectiveness factor
VD = overall effectiveness factor for a system completely
dominated by external diffusion
qs = effectiveness factor related to the catalyst particles
X = anthraquinone conversion
p , = density of catalyst particle (g/cm3)
4 = generalized Thiele modulus
PL

Literature Cited
Akita, K.; Yoshida, F. Ind. Eng. Chem. Process Des. Dev. 1973,12,
76.
Berglin, T.; ShiiGn, N. H. Ind. Eng. Chem. Process Des. Dev. 1981,
20, 615.
Berglin, T.; Shiiiin, N. T. Ind. Eng. Chem. Process Des. Dev. 1983,
22, 150.
Chaudari, R. V.; Ramachandran, P. A. Ing. Eng. Chem. Fundam.
1980, 19, 201.
Kirdin, K. K.; Franchuk, V. I.; Balabin, I. Y.; Aaanova, M. I. Sou.
Chem. Znd. 1970,9,5.
Kirk, T.; Othmer, K Encyclopedia of Chemical Technology;Wiley:
New York, 1981; Vol. 13.
Link, V.; Vacek, Vi Chem. Eng. Sci. 1981,36,1741.
Norris, A. C. Computational Chemistry; John Wiley: New York,
1981.
Powell, R. Hydrogen Peroxide; Noyes: New York, 1968.
Ramachandran, P. A.; Chaudari, R. V. Three Phase Catalytic
Reactors; Gordon and Breach London, 1983.
Sano, Y.; Yamaguchi, N.; Adachi, T. J . Chem. Eng. Jpn. 1974,7,255.
Santacesaria, E.; Wilkinson, P.; Babini, P.; Carrl, S. Znd. Eng. Chem.
Res. 1988,27 (5),780.
UlmannsEncyclopedia of Industrial Chemistry;VCH Weinheim,
1989; Vol. A13, p 443.
Wilke, C. R.; Chang, P. AIChE J . 1965, 1, 264.

Received for review July 8, 1993


Revised manuscript received September 20, 1993
Accepted November 2 , 1993.
@

Abstract published in Advance ACS Abstracts, January 1,

1994.

Вам также может понравиться