You are on page 1of 10

+ +

1496 Ind. Eng. Chem. Res. 1996, 35, 1496-1505

Kinetics of Front-End Acetylene Hydrogenation in Ethylene


Production
Noemı́ S. Schbib, Miguel A. Garcı́a, Carlos E. Gı́gola, and Alberto F. Errazu*
Planta Piloto de Ingenierı́a Quı́mica, UNS-CONICET, 12 de octubre 1842, 8000 Bahı́a Blanca, Argentina

The kinetics of acetylene hydrogenation in the presence of a large excess of ethylene was studied
in a laboratory flow reactor. Experiments were carried out using a Pd/R-Al2O3 commercial
catalyst and a simulated cracker gas mixture (H2/C2H2 ) 50; 60% C2H4; 30% H2, and traces of
CO), at varying temperature (293-393 K) and pressure (2-35 atm). Competing mechanisms
for acetylene and ethylene hydrogenation were formulated and the corresponding kinetic
equations derived by rate-determining step methods. A criterion based upon statistical analysis
Downloaded via VIRGINIA POLYTECH INST STATE UNIV on November 10, 2021 at 17:15:05 (UTC).

was used to discriminate between rival kinetic models. The selected equations are consistent
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

with the adsorption of C2H2 and C2H4 in the same active sites followed by reaction with adsorbed
hydrogen atoms to form C2H4 and C2H6 in a one-step process. Good agreement between computed
and experimental results was obtained using a nonisothermal reactor model that takes into
account the existence of external temperature and concentration gradients. The derived kinetic
equations together with a pseudohomogeneous model of an integral adiabatic flow reactor were
employed to simulate the conversion and the temperature profiles for a commercial hydrogenation
unit.

Introduction the hydrogenation of C2H2 in the presence of C2H4.


Most studies have been performed with pure C2H2
The selective hydrogenation of acetylene in the pres- under conditions far removed from industrial opera-
ence of large amounts of ethylene is a process of tions. Some papers deal with the hydrogenation of
considerable importance in the manufacture of polymer- C2H2-C2H4 mixtures (McGown et al., 1978; Moses et
grade ethylene. Small amounts of alkyne, in the parts al., 1984; Adúriz et al., 1990). Orders in H2 lie between
per million range, affect the polymerization catalyst. 1 and 1.6, and that for C2H2 has been found to be zero
Thus the acetylene concentration in the product stream or negative. Activation energy values between 9 and
must be reduced to 5 ppm or less. 16 kcal/mol have been quoted. For C2H4 hydrogenation
on Pd the information is more scarce. Schuit and Van
This hydrogenation process could be located at dif-
Reijen (1958) reported an activation energy of 8.4 kcal/
ferent points in the purification section of an ethylene
mol, a H2 order of 0.6, and an C2H4 order of 0, for Pd on
plant (Lam and Lloyd, 1972; Derrien, 1986). In one
SiO2. Weiss et al. (1984) have investigated the effect of
scheme the converter is placed after the conversion
CO on C2H2 and C2H4 conversion. The former was
section, following a caustic scrubbing treatment to
found to be 50-60 times lower in the presence of CO
eliminate CO2. Another alternative involves the hy-
(800 ppm) than in its absence, but the effect was small
drogenation of C2H2 in the C2H4-rich stream taken from
at low H2 concentration. On the other hand, the
the top of the de-ethanizer. The first alternative is
hydrogenation of C2H4 was almost suppressed by the
known as front-end hydrogenation and the second one
addition of CO.
as tail-end hydrogenation.
The tail-end hydrogenation process, under conditions
Quite different operating conditions are used in each
similar to those of industrial reactors, has been studied
case. The front-end configuration involves the hydro-
in a pilot plant by Battiston et al. (1982). The experi-
genation of C2H2 in the raw cracked-gas mixture so that
mental results were correlated by an empirical model
the stream has a high H2/C2H2 (≈100) ratio and several
to take into account the influence of the several reaction
acetylenic, olefinic, and diolefinic components. Carbon
variables on C2H2. Increasing the H2/C2H2 ratio, at
monoxide also appears in the feed due to the inverse
constant CO concentration, increases the conversion of
water-gas shift reaction in the cracking furnaces. They
both C2H2 and C2H4. A more standard approach was
all act as inhibitors of C2H4 hydrogenation. Pure C2H2
used by Men’shchikov et al. (1975). A commercial Pd
in the tail-end converter is dosed with H2 to obtain a
catalyst was tested in a flow reactor at 20 atm, in the
stoichiometric H2/C2H2 ratio.
353-430 K temperature range using a tail-end mixture.
Supported palladium catalysts with low metal content Rate equations for C2H2 and C2H4 hydrogenation, which
are used for both processes due to their exceptional high were written under the assumption that the reactions
activity and selectivity. It is important to point out that occur on separate sites, gave good correlation of experi-
Pd is unique, among other metals, for the preferential mental data.
hydrogenation of alkynes and diolefins in the presence
Studies of this kind for front-end C2H2 hydrogenation
of olefins. However the C2H4 f C2H6 reaction becomes
have not been reported. In this paper, laboratory data
significant when the C2H2 concentration is low, that is,
obtained on a wide range of pressure, temperature, and
at high levels of conversion.
conversion values were fitted to several kinetic equa-
Despite the commercial importance of C2H4 purifica- tions for C2H2 and C2H4 hydrogenation. The selected
tion processes, there is limited kinetic information on models were used to simulate the operation of an
industrial acetylene converter. Predicted conversion
* Author to whom correspondence should be addressed. and temperature profiles were compared with plant
E-mail: reschbib@criba.edu.ar. FAX: 54-91-883764. data.
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1497


Methods

(6) ES + 2HS f EtS + 2S


(3) AS + 2HS f ES + 2S
In the present study the experimental and theoretical

(7) ES + H2(g) f EtS


(4) AS + H2(g) f ES
work has been organized as follows:

(2) H2 + 2S f 2HS

(9) CO + S f COS
scheme 7

(8) Et + S f EtS
1. Several reaction schemes were written for C2H2

(5) E + S f ES
(1) A + S f AS
and C2H4 hydrogenation on Pd using information from
fundamental kinetic studies.
2. The rate-determining step method was applied to
the proposed mechanisms in order to derive the corre-
sponding kinetic equations.
3. Conversion versus temperature curves for C2H2

(8) EHS + HS f EtS + S


and C2H4 hydrogenation were obtained in a laboratory

(3) AS + HS f AHS + S
(4) AHS + HS f ES + S

(7) ES + HS f EHS + S
reactor, in the 1-30 atm pressure range, using a

(9) ES + H2(g) f EtS


(5) AS + H2(g) f ES

(11) CO + S f COS
(2) H2 + 2S f 2HS
commercial Pd/R-Al2O3 catalyst.

(10) Et + S f EtS
scheme 6
4. The rate equations were coupled to three reactor

(6) E + S f ES
(1) A + S f AS
models of increasing complexity. The algebraic and
differential equations, which are not linear in the kinetic
parameters, define a set of regression models.
5. A nonlinear regression routine that selects the
conversion as a response variable was used to estimate
the unknown parameters.

(7) EHS + HZ f EtS + Z


(3) AS + HZ f AHS + Z
(4) AHS + HZ f ES + Z

(6) ES + HZ f EHS + Z
6. The discrimination between rival models was
based upon the requirement of positive parameters

(2) H2 + 2Z f 2HZ

(9) CO + S f COS
supplemented with a variance analysis performed on

scheme 5

(8) Et + S f EtS
the conversion residuals.

(5) E + S f ES
(1) A + S f AS
Kinetic Models
The first task of the present study was to derive
kinetic equations that could be compared with experi-
mental data for parameter estimation and discrimina-

(5) ES + 2HZ f EtS + 2Z


(3) AS + 2HZ f ES + 2Z
tion. Several sequences of elementary steps were
written on the basis of information available from
previous studies. It is well-known that H2 is dissocia-

(2) H2 + 2Z f 2HZ

(7) CO + S f COS
scheme 4

(6) Et + S f EtS
tively adsorbed on Pd to form H-Pd surface entities.

(4) E + S f ES
(1) A + S f AS
On the other hand, it is accepted that C2H2 and C2H4
adsorption occurs by an associative mechanism involv-
Table 1. Reaction Mechanism for the Selective Hydrogenation of C2H2 and C2H4

ing one or two Pd surface atoms (π-bonded or σ-bonded


structures). Consequently it was considered that simi-
lar or different types of sites may exist for H2 and
hydrocarbon adsorption. For the interaction of CO with
(4) H2(g) + ES f EtS

the metal surface, an associative mechanism involving


(2) H2(g) + AS f ES

(6) CO + S f COS

hydrogen adsorption sites has been postulated.


(5) Et + S f EtS
scheme 3

(3) E + S f ES
(1) A + S f AS

Regarding the surface reactions, three different reac-


tion paths were assumed: a stepwise addition of ad-
sorbed hydrogen atoms, the simultaneous addition of
two atoms, or the reaction of gaseous hydrogen with
adsorbed hydrocarbons. It has already been shown
(Adúriz et al., 1990; Sarkany et al., 1986) that C2H6
formation during hydrogenation of C2H2-C2H4 mixtures
(5) ES + 2HS f EtS + 2S
(3) AS + 2HS f ES + 2S

is mainly due to C2H4 hydrogenation. Consequently we


have excluded the direct formation of C2H6 from C2H2.
(2) H2 + 2S f 2HS

(7) CO + S f COS

Reaction schemes 1-5, shown in Table 1, were written


scheme 2

(6) Et + S f EtS

on the basis of these considerations. In principle all


(1) A + S f AS

(4) E + S f ES

steps were assumed to be reversible. If the stepwise


addition of adsorbed hydrogen and the Eley-Rideal
mechanism operate simultaneously (schemes 1 and 3,
respectively), scheme 6 is obtained. On the other hand,,
the combination of schemes 2 and 3 leads to scheme 7.
In order to obtain the steady-state kinetic models the
(7) EHS + HS f EtS + S
(3) AS + HS f AHS + S

(6) ES + HS f EHS + S
(4) AHS + HS f ES + S

rate-determining-step method was used. In principle


any step can be rate limiting, but it is possible to reduce
(2) H2 + 2S f 2HS

(9) CO + S f COS

the number of equations on the basis of information


scheme 1

(8) Et + S f EtS

available from surface chemistry studies. Hydrogen


(5) E + S f ES
(1) A + S f AS

adsorption is a fast, nonactivated process, and therefore


this step could be assumed to be at equilibrium. The
desorption of ethane cannot be the rate-limiting step,
for ethylene hydrogenation, because it is readily dis-
placed from the surface by the strongly bound hydro-
+ +

1498 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Table 2. Kinetic Equations Derived from the Reaction Schemes Presented in Table 1a
r2,14 ) r7,15 rA ) kACA/D
rE ) kECE/D
D ) 1 + (KHCH)1/2 + CEt ∑i)1
2
KiCH-i + KEtCEt + KCOCCO
r1,36 1/2
rA ) kACACH /D 2

rE ) kECECH1/2/D2
D ) 1 + KACA + KECE + (KHCH)1/2 + (K1CE + K2CEt)CH-1/2 + KEtCEt + KCOCCO
r1,47 rA ) kACACH/D2
rE ) kECECH/D2
D ) 1 + KACA + KECE + (KH + K1CA + K2CE)CH1/2 + KEtCEt + KCOCCO
r2,35 rA ) kACACH/D3
rE ) kECECH/D3
D ) 1 + KACA + KECE + (KHCH)1/2 + KEtCEt + KCOCCO
r3,13 ) r4,14 rA ) kACA/D
rE ) kECE/D
D ) 1 + (KEt + K1CH-1 + K2CH-2)CEt + KCOCCO
r3,24 rA ) kACACH/D
rE ) kECECH/D
D ) 1 + KACA + KECE + KEtCEt + KCOCCO
r4,35 rA ) kACACH/D
rE ) kECECH/D
D ) (1 + KACA + KECE + KEtCEt + KCOCCO)(1 + KH1/2CH1/2)2
r5,15 rA ) kACA/D
rE ) kECE/D
D ) 1 + (KEt + K1CH-1/2 + K2CH + K3CH-3/2 + K4CH-2)CEt + KCOCCO
r5,36 rA ) kACACH1/2/D
rE ) kECECH1/2/D
D ) (1 + KACA + KECE + (K1CE + K2CEt)CH-1/2 + KEtCEt + KCOCCO)(1 + KH1/2CH1/2)
r5,47 rA ) kACACH/D
rE ) kECECH/D
D ) (1 + KACA + KECE + (K1CA + K2CE)CH1/2 + KEtCEt + KCOCCO)(1 + KH1/2CH1/2)
r6,16 ) r1,15 rA ) kACA/D
rE ) kECE/D
D ) 1 + (KHCH)1/2 + (K1CH-1/2 + K2CH-1 + K3CH-3/2 + K4CH-2)CEt + KEtCEt + KCOCCO
r6,35/79b rA ) kACACH1/2/D2 + k*ACACH/D
rE ) kECECH1/2/D2 + k*ECECH/D
D ) 1 + KACA + (KHCH)1/2 + KECE + (K1CE + K2CEt)CH-1/2 + KEtCEt + KCOCCO
r6,45/89b rA ) kACACH/D2 + k*ACACH/D
rE ) kECECH/D2 + k*ECECH/D
D ) 1 + KACA + KECE + (KH1/2 + K1CA + K2CE) CH-1/2 + KETCET + KCOCCO
r7,34/67b rA ) kACACH/D3 + k*ACACH/D
rE ) kECECH/D3 + k*ECECH/D
D ) 1 + KACA + (KHCH)1/2 + KECE + KEtCEt + KCOCCO
a The first digit identifies the sequence of steps and the others the controlling steps for C H and C H hydrogenation. b Simultaneous
2 2 2 4
mechanisms with two different controlling steps for acetylene and ethylene hydrogenation are considered.

carbons (ethylene and acetylene). In addition it has equations the following constants were introduced
been shown recently (Park and Price, 1991) that CO
KEt KEt
enhances the desorption of C2H4. Consequently only the K1 ) K2 ) (2)
adsorption of C2H2 and C2H4 and the surface interac- KHKES 2
KH KESKAS
tions were chosen as rate-controlling steps to derive the
kinetic models. To further reduce the number of equa- Using these constants, [S] may be written in terms
tions, the same determining step was selected for C2H2 of the gas phase concentrations
and C2H4 hydrogenation. By use of this approach 14 [S] ) [ST]/D
kinetic models were derived, as shown in Table 2.
As an example let us consider scheme 2 when the where
adsorption of C2H2 and C2H4, steps 1 and 4, are rate- 2


controlling. The corresponding kinetic equations are
D ) 1 + (KHCH)1/2 + CET KiCH-i +
i)1
r2,1 ) -rA ) k1CA[S]
KETCET + KCOCCO (3)
r2,4 ) -rE ) k4CE[S] (1) and [St] is the concentration of free and occupied sites.
Combining the previous equations, the rates of disap-
where k1 and k4 are the rate constants and [S] is the pearance of C2H2 and C2H4 are
concentration of free surface sites. Considering the -rA ) kACA/D and -rE ) kECE/D (4)
irreversible nature of the overall hydrogenation reac-
tions, the reverse reactions of steps 1 and 4 have been where
neglected. The remaining steps, 2, 3, 5, 6, and 7 are
considered to be in a quasiequilibrium state. We use kA ) k1[St] and kE ) k4[St]
letters to identify the adsorption-desorption processes Experimental Section
at equilibrium: KH for H2, KEt for C2H6 , KE for C2H4,
and KCO for CO, while the constants KAS and KES take A laboratory scale reactor was used to obtain conver-
into account the surface reactions of adsorbed C2H2 and sion vs temperature curves in the 1-30 atm pressure
C2H4 at equilibrium. In order to simplify the rate range. Catalyst charges of 1-3 g were packed in a
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1499


Table 3. Feed Mixture and Operating Conditions for the
Laboratory Reactor
components composition (mol % ( 0.02)
CH4 6.53
C2H2 0.52
C2H4 43.99
C2H6 26.72
H2 22.1
CO 150-5000 ppm
flow rate 100 cm3 (STP)/min
mass of catalyst 1-3 g
pressure 1-35 atm
temperature 288-393 K

Table 4. Catalyst Characterization


support R-Al2O3 Figure 1. Acetylene and ethylene conversion as a function of
palladium content (wt %) 0.04 temperature. Feed mixture as given in Table 3. SV ) 33 cm3 gcat-1
BET surface area (m2/g) 21 min-1. CCO ) 800 ppm. Pressure: 0, 1 atm; O, 9 atm.
metal dispersion (H/Pd) 0.15
apparent density (g/cm3) 1.61 Table 5. Effect of Pressure on C2H2 and C2H4 Conversion
real density (g/cm3) 3.61 (CO Content ) 1400 ppm; SV ) 100 cm3 STP gcat-1 min-1)
average pore diameter (Å) 1000
pore volume (cm3/g) 0.338 conversion (%)
porosity (%) 54.5 temp (K) 1 atm 8 atm 16.50 atm 32 atm
348 C2H2 85.54 95.86 95.39 94.13
stainless-steel tube (10 mm i.d.) immersed in a dieth- C2H4 0.00 0.00 0.00 0.00
ylene glycol bath. The total pressure was adjusted by 359 C2H2 94.44 97.75 97.75 96.64
means of a back-pressure regulator. Experiments were C2H4 0.00 1.02 1.67 4.50
performed in the 288-393 K temperature range while 368 C2H2 96.13 98.16 98.02 96.97
the gas flow rate was held constant at 100 cm3 STP C2H4 0.00 3.12 6.07 14.50
min-1 using a mass flow controller. A thermocouple in Table 6. Effect of CO on C2H2 and C2H4 Conversion at
contact with the catalyst particles was located in an Constant Pressure (Pressure ) 16 atm; SV ) 100 cm3 STP
axial position to monitor the reaction temperature. In gcat-1 min-1)
addition the wall temperature was measured. conversion (%)
The analysis of reactants and products was performed
by on-line gas chromatography using a Carbosieve S-II CO content (ppm) temp (K) C2H2 C2H4
column (1/8 in. × 80 cm) held at 393 K. Fresh catalyst 150 341 98.31 10.45
samples were pretreated in N2 at 433 K for 8 h before 1400 359 98.99 6.70
feeding the hydrocarbon mixture. 5000 393 98.62 6.08
The reaction mixture was prepared by mixing CP- ) 1%. Consequently the Pd/R-Al2O3 catalyst, under the
grade hydrocarbons with high-purity hydrogen and CO. present experimental conditions, is very selective in the
Table 3 summarizes the gas composition and the main sense that it allows the total elimination of C2H2 with
operating conditions. minimum C2H4 losses. Increasing the total pressure
A commercial catalyst, ICI 38-1, containing 0.04% accelerates both reactions, and therefore similar conver-
Pd, was used in this study. The main catalyst proper- sions are obtained at lower temperatures. However a
ties are shown in Table 4. more detailed testing, in the region of high C2H2
The effect of increasing temperature on C2H2 and conversion, indicates that the effect of pressure is more
C2H4 conversion was investigated in the presence of marked on the rate of C2H4 hydrogenation. Table 5
varying amounts of CO and a total pressure of 1, 9, 16, shows that the conversion of C2H2, at constant temper-
and 32 atm. In these runs the space velocity was held ature, increases with pressure in the 1-9 atm range
constant. When the conversion of C2H2 was almost but remains constant afterward. On the other hand,
100%, the heat of reaction produced a small temperature the hydrogenation of C2H4 reflects a steady increase in
difference between the catalyst bed and the reactor wall. conversion from 1 to 32 atm. Consequently the tem-
The maximum C2H4 conversion was about 10-15%. A perature span between complete elimination of C2H2
large experimental error was present in this measure- and the onset of C2H4 hydrogenation is reduced and the
ment as compared with C2H2 conversion, due to the high catalyst becomes less selective.
concentration of C2H4 and C2H6 in the feed mixture. At The effect of CO on activity and selectivity, for
high C2H4 conversion isothermal conditions could not experiments performed at 16 atm, is shown in Table 6.
be maintained and a large temperature gradient was Increasing the concentration of CO, from 150 to 1400
detected by the thermocouples. In some experiments ppm, lowered the C2H2 hydrogenation rate, and a higher
the temperature was held constant and the conversion temperature was needed to obtain nearly the same
was measured as a function of the total pressure. conversion. Despite the large increase in temperature,
The effect of temperature on C2H2 and C2H4 conver- 18 K, the conversion of C2H4 decreased from 10.85% to
sion, at 1 and 9 atm and 800 ppm CO, is shown in 6.70%. When the CO concentration was raised to 5000
Figure 1. The initial trend is a large increase in C2H2 ppm, a further increase in temperature was required
conversion within a narrow temperature range, followed to maintain the same conversion of C2H2. However the
by a marked decrease in the rate of reaction above 90% conversion of C2H4 was almost the same: 6.08%. These
conversion. In this case the conversion of ethylene was results indicate that the concentration of CO at >150
observed after the complete elimination of C2H2, with ppm affects the rate of C2H2 and C2H4 hydrogenation
a temperature span of about 15 K from χA ) 100% to χE to the same extent.
+ +

1500 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Reactor Models Kutta method. In this way, the C2H2 and C2H4 conver-
sions (〈χi〉) at the reactor outlet are obtained.
In order to process the experimental data, a suitable A more complex reactor model (model III) emerges if
reactor model must be used. As a first approximation, one assumes the existence of temperature and concen-
a perfectly mixed flow reactor was considered (model tration gradients in the gas phase region adjacent to
I). Although this model is not appropriate to describe the catalyst particles. Therefore T * Ts and Ci * Csi.
our laboratory setup, it has been chosen in an attempt This situation is likely to occur in the laboratory reactor
to evaluate a large number of kinetic equations without due to the low gas velocity over the catalyst particles.
considerable computing effort. The basic mass balance Under steady-state conditions the rate of mass trans-
equations were written as port of a reactant (i) through the gas phase film is equal
to the observable surface reaction rate (rsi):
w χi
(5)
Fi ri
)
rsi ) kmiam(Ci - Csi) (12)

where i refers to C2H2 and C2H4. They can be readily Introducing the external effectiveness factor ηi,
coupled to the rate models given in Table 2 to relate
the fractional conversion χi to the kinetic parameters. rsi ) ηirio (13)
Therefore a set of nonlinear algebraic equations was
obtained. Using experimental w/Fi and T values as where rio is the maximum rate or reaction obtained
independent variables and assuming tentative kinetic when Csi ) Ci. In addition the Damkohler number Dio
constants, the conversion (〈χi〉) was estimated and is defined as the ratio of the maximum surface rate to
compared with the measured values (χi). the maximum mass transport rate:
The second reactor model (model II) was assumed to
be nonisothermal, isobaric with a plug flow pattern, and Dio ) rio/(kmiamCi) (14)
pseudohomogeneous. In this case the mass and energy
balance defined a set of nonlinear ordinary differential Following the treatment of Carberry (Carberry, 1976)
equations coupled through the rate model ri: eqs 13 and 14 can be related to define a dimensionless
observable quantity:
d(vCi)
) (-ri)Fcat (6)
dz ηiDio ) rsi/(kmiamCi) (15)
dT
∑i (-∆Hi)(-ri) - 4U(T - Tw)
Combining eqs 12 and 15, a dimensionless concentra-
cp ) Fcat (7) tion in terms of an observable quantity may be obtained:
dz
C*
i ) Csi/Ci ) 1 - ηiDio
The overall heat transfer coefficient U was defined
as follows:
For C2H2,
1
(ln(dto/dt)
)
-1
U) (8) C*A ) CsA/CA ) 1 - ηADAo (16)
hA 2FkwA
+

Taking into account that ethylene is a product and


In this equation the heat transfer coefficient h takes also a reactant, the concentration on the gas-solid
into account the total resistance to heat flow in the interphase is given by
radial direction and it is given by the Crider and Foss
equations (Crider and Foss, 1965). C*E ) 1 - ηEDEo + ηADAokmACA/(kmECE) (17)
1/h ) 1/hw + dt/6kref (9) On the other hand, the H2 concentration is related to
the consumption of both C2H2 and C2H4:
hw ) 3krefRe -0.25
/dt (10)
C*H ) 1 - ηADAokmACA/(kmHCH) -
where hw is the heat transfer coefficient at the tube wall. ηEDEokmECE/(kmHCH) (18)
The radial effective thermal conductivity kref was ob-
tained from the Dixon and Cresswell correlation (Dixon
Under steady-state, nonisothermal conditions, the

[ ]
and Cresswell, 1979)
surface temperature could be related to the bulk gas
phase temperature through an energy balance on the
1 + (8krf/hwfdt)
catalyst particles.

( )
kref ) krf + krs (11)
1 0.1
(16/3krs)
hfsdp kp hfsam(Ts - T) ) (-∆HA)(-rA) + (-∆HE)(-rE) (19)
+
1+ 2
(1 - )(dt/dp) Dividing this equation by (kmHamT) and recalling the
jD ) jH heat and mass transfer analogy,
where krf and krs are the radial conductivities of the fluid
and the solid, respectively. The effective conductivity Fcp(Sc/Pr)2/3(T* - 1) ) (-∆HA)(-rA)/(kmHamT) +
depends also on the fluid/wall (hwf) and the fluid/solid
(hfs) heat transfer coefficients. Suitable correlations for (-∆HE)(-rE)/(kmHamT)
these coefficients are found in the previous reference. T* ) 1 + [(-∆HA)(-rA)/(kmHamTFcp)](Pr/Sc)2/3 +
The previous equations constitute an initial value
problem that has been solved by means of a Runge- [(-∆HE)(-rE)/(kmHamTFcp)](Pr/Sc)2/3 (20)
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1501


Using eq 15 the dimensionless temperature may be eter, and it continues until the variance analysis
written indicates that a reasonable agreement between experi-
mental and calculated conversions has been found. A
T* ) 1 + βAηADAo + βEηEDEo (21) set of 163 experimental points was used for these
estimations.
where For the more complex reactor models, the procedure
described above uses a large computing time due to the
βi ) [(-∆Hi)(kmiCi)/(kmHTFcp)](Pr/Sc)2/3 i ) A, E fact that the calculated conversion is obtained by
numerical integration of differential equations. In
The temperature dependence of the rate constant ki and addition each step of the regression routine requires
the adsorption equilibrium constant Ki may also be about 2600-3200 simulations to evaluate the Jacobian.
written in a dimensionless form: In order to reduce computing time, a shortcut method
was adopted, based on sequential discrimination, so that
k*i ) ksi/ki (or K*
i ) Ksi/Ki) ) exp(-E*
i (1/T* - 1)) many rate models were eliminated early in the process
(22) using model I, where only one algebraic equation needs
to be solved. A set of parameters was estimated for each
rate equation. The variance of the experimental and
where: calculated conversion values provided a convenient
E* method to analyze the quality of this fitting procedure.
i ) Ei/RT
In addition the discrimination between the different rate
To integrate eqs 6 and 7 for a given axial coordinate, equations was done on the basis of negative parameters
eqs 16-22 must be solved to find the concentrations and or unexpected trends in the conversion versus temper-
temperature at the catalyst surface. Consequently, a ature curves. In this way 11 rate models were elimi-
system of coupled nonlinear differential and algebraic nated, as shown in Table 7.
equations has to be solved. In this work, the ordinary In the following stage the remaining kinetic equations
differential equations were integrated with the Runge- were processed with the same regression routine but
Kutta-Gill method and the algebraic equations were using model II. The parameters derived in the previous
solved using a quasi-Newton algorithm, where the first stage were now used as initial values. One kinetic
Jacobian was numerically evaluated and then updated model was eliminated when the discrimination criterion
by means of Broyden’s method (Paloschi and Perkins, was applied. Finally the parameters determined with
1988) at each iteration step. Thus the overall computing model II were adopted as initial values to fit the
time was greatly reduced. experimental data with model III. Therefore the adopted
A nonlinear regression routine based on the Mar- procedure uses a simple reactor model when the solution
quardt algorithm was used for the three reactor models is a long way off, and the more realistic but time-
described above to adjust the kinetic parameters cor- consuming models were applied when few iterations
responding to the rate models given in Table 2. The were needed.
objective function to be minimized was defined as The Kinetic Equations. The best fit with the
experimental data was provided by the rate model r2,35
φ) ∑i [χi - 〈χi〉]2 (23)
that assumes the adsorption of C2H2, C2H4, and H2 on
the same type of sites and the simultaneous addition of
two hydrogen atoms to the absorbed hydrocarbons as
In the models discussed above we have neglected the determining steps. Table 8 presents the selected kinetic
presence of internal diffusion limitations. This simpli- expressions and the parameter values obtained by
fying assumption was based on the fact that the average minimization of the sum of residual squares. Upon
pore diameter of the catalyst pellets was quite large, examination these equations reveal some interesting
1000 Å. However, due to the high hydrogenation rate features. First it is observed that the hydrocarbon
and the homogeneous distribution of Pd in the pellet, adsorption constants KE, KET, and KA have a negligible
internal concentration gradients may be present. In effect on the rates and consequently the kinetic equa-
order to check for the absence of intrapellet diffusion tions can be simplified to
limitations, we selected the criterion of Weisz and Prater kACACH
(Froment and Bischoff, 1990) which in turn requires a -rA ) (24)
measured value of the rate of reaction. Consequently [1 + (KHCH)1/2 + KCOCCO]3
application of this criterion was performed after ad-
equate kinetic equations were available to calculate the kECECH
rates of C2H2 and C2H4 hydrogenation at different -rE ) (25)
[1 + (KHCH)1/2 + KCOCCO]3
conversion levels.

Results and Discussion It is observed that the rate of C2H4 hydrogenation


does not depend on the concentration of C2H2 as
Determination of Kinetic Constants and Model previously reported by Margitfalvy et al. (1981). The
Discrimination. The process parameters were con- explanation may be found in the presence of CO. As
sidered as independent or input variables to be intro- pointed out by Bos and Westerterp (1993a), CO may
duced as data in the mathematical models. On the suppress the influence of C2H2 in C2H4 hydrogenation.
other hand, the measured conversion values were the Another surprising feature is the positive values for
dependent variables to be compared with the theoretical the adsorption energy on the CO and H2 constants
predictions. From this comparison the regression rou- which indicates that they will increase with tempera-
tine estimates the kinetic parameters. The rate equa- ture. It is reasonable to expect that the adsorption
tions of Table 2 were first included in model I. The terms decrease with temperature due to the exothermic
calculation starts with an initial guess for each param- character of the adsorption processes. This anomalous
+ +

1502 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

Table 7. Discrimination Sequence for the Kinetic Models of Table 2


model I model II model III
selection selection selection
rii procedure variance procedure variance procedure variance
r2,14 ) r7,15 a 0.366E-1
r1,36 b 0.481E-2 c 0.307E-2
r1,47 a 0.402E-2
r2,35 b 0.487E-2 b 0.270E-2 b 0.229E-2
r3,13 ) r4,14 c 0.319
r3,24 a 0.127E-1
r4,35 a 0.426E-1
r5,15 c 0.838E-2
r5,36 a 0.662E-2
r5,47 a 0.211E-1
r6,16 ) r1,15 a 0.592E-2
r6,35/79 a 0.779E-2
r6,45/89 a 0.340E-2
r7,34/67 b 0.273E-2 b 0.239E-2 c 0.819E-2
a Adjusted by regression. b Rejected because Eact < 0.b c Rejected because of high variance.

Table 8. Selected Rate Equation and Kinetic pensated by the increase in temperature, so the rate is
Parameters for C2H2 and C2H4 Hydrogenation nearly constant in the 325-340 K temperature range.
kinetic parameters log Ai E (kcal/gmol) Eventually the decrease in C2H2 becomes so important
kA (m6/(s gcat kmol)) 31.1 45.51 that the rate falls below that of C2H4, so at high
kE (m6/(s gcat kmol)) 26.6 42.94 conversion values there is a marked decrease in selec-
KH (m3/kmol) 20.2 21.22 tivity.
KCO (m3/kmol) 13.6 9.95 It should be noted that our kinetic equations do not
KE (m3/kmol) 0.26 0.005 predict a change in selectivity with the CO content
KEt (m3/kmol) 0.001
-0.012
because the rA/rE ratio depends only on the concentra-
KA (m3/kmol) 0.001
tion of C2H2 and C2H4. In other words, the presence of
-16

kACACH CO seems to affect the rate of C2H2 and C2H4 hydroge-


-rA )
[1 + (KHCH) 1/2
+ KACA + KECE + KEtCEt + KCOCCO]3
nation to the same extent, in accordance with the results
of Table 6. As the concentration of CO increases, a
kECECH higher temperature is required to obtain a given conver-
-rE ) 1/2 sion of C2H2, but the conversion of C2H4 remains the
[1 + (KHCH) + KACA + KECE + KEtCEt + KCOCCO]3
same. The effect of CO may be explained by assuming
where kA ) k3St3KAKH and kE ) k5St3KAKH that it limits the amount of adsorbed H2 in accordance
with kinetic models based on the adsorption of H2 and
result may be due to data error, inadequate selection CO on the same type of sites. However it is important
of experimental conditions, or data analysis. Problems to stress that the presence of a minimum level of CO is
of this kind are often encountered on rate modeling as essential to avoid the uncontrolled hydrogenation of
discussed by Doraiswamy and Sharma (1984). In spite C2H4.
of this difficulty the hydrogenation rates predicted by As mentioned by one reviewer, there is evidence of
our models exhibit a normal Arrhenius behavior because adsorption of acetylene and ethylene on separate sites,
the effect of temperature on the adsorption parameters which may suggest additional hydrogenation mecha-
is overcompensated by the large activation energy on nisms. We are aware of several studies that postulated
the rate constants kA and kE. different sites for acetylene and ethylene hydrogena-
Equations 24 and 25 and those of model III were used tiona (All-Ammar and Webb, 1978; McGown et al., 1978;
in the simulation of an adiabatic reactor to observe the Leviness et al., 1984). Therefore a reaction scheme
predicted rate versus temperature profiles, as shown in based on a two-site mechanism could have been included
Figure 2. The rate of C2H2 hydrogenation is initially in Table 1. However this model would lead to more
larger than that of C2H4, as expected. As the reaction complex kinetic equations as shown by Gva and Kuo
proceeds, the decrease in C2H2 concentration is com- (1988), increasing the number of adjustable parameters.
In addition, our experimental results indicate that the
presence of CO prevents the hydrogenation of ethylene
up to a very high level of acetylene conversion, as
observed in Figure 1. Consequently, a two-site mech-
anism that may be relevant under different experimen-
tal conditions was not used here to obtain additional
kinetic equations.
Quite recently Bos et al. (1993b) used rate models
from previous studies (Men’shchikov et al., 1975; Gva
and Kuo, 1988) to correlate kinetic data obtained with
a typical tail-end hydrogenation mixture. The original
equations were based on the assumption that the
hydrogenation of C2H2 and C2H4 proceeds on different
Figure 2. Acetylene and ethylene hydrogenation rate vs tem-
types of sites, and they were properly modified to take
perature for an adiabatic reactor. Kinetic equations of Table 8. into account the effect of CO. In this case the temper-
Feed mixture as given in Table 3. T0 ) 325 K, P ) 16.3 atm; CCO ature dependence of the adsorption terms was according
) 1400 ppm. to expectations; that is they decrease with an increase
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1503


near the reactor entrance. A sharp rise in temperature
profiles is followed by a slow decrease along the axial
coordinate. The temperature overshoot depends on the
feed-reactor wall temperature: at 325 K, with a conver-
sion of 70.7% it is about 1.5 K. Increasing the inlet
temperature to 345 K, the C2H2 conversion was close
to 98% and the temperature overshoot exceeded 3 K.
The simulations also indicate that a temperature gradi-
ent of about 0.5 K, between the exit gas and the reactor
wall, should be expected. As mentioned in the Experi-
mental Section, temperature differences of this magni-
tude were observed in our runs, which shows the
convenience of using a nonisothermal model to fit the
Figure 3. Calculated conversion and temperature profiles for the
laboratory reactor using reactor model III and the rate equations
experimental data.
of Table 8. Feed mixture as given in Table 3 T0 ) 325 K; P ) 16.3 As described above, reactor model III takes into
atm; CCO ) 1400 ppm. 9, Experimental data. account the presence of temperature and concentration
differences between the catalyst particle and the gas
phase. The calculated concentration gradients were
found to be negligible for all runs. On the other hand,
the ∆T ) Ts - T values were found to depend on the
gas phase temperature which in turn is related to the
rate of reaction. At the reactor entrance, in the hot-
spot region, large temperature gradients between the
gas phase and the catalyst particles were calculated. In
Figure 3, at 326 K the ∆T value was about 8 K. It
increased to nearly 20 K when the gas phase temper-
ature overshot to 348 K, as was the case in Figure 4.
These results confirm the suitability of reactor model
III to treat the laboratory data in order to obtain the
kinetic parameters.
Figure 4. Calculated conversion and temperature profile for the Industrial Reactor Simulation. In order to check
laboratory reactor using reactor model III and the rate equations the validity of our kinetic equations to predict conver-
of Table 8. Feed mixture as given in Table 3, T0 ) 345 K; P ) sion and temperature profiles, an attempt was made to
16.3 atm.; CCO ) 1400 ppm. 9, Experimental data. simulate the operation of an industrial acetylene hy-
in temperature. The equations of Bos et al. predict a drogenation reactor. The process scheme consists of an
larger difference between acetylene and ethylene hy- ethane cracker followed by three adiabatic reactors in
drogenation rates due to the reduced partial pressure series in a typical front-end hydrogenation configura-
of hydrogen. They also found that the addition of CO tion. The simulation was restricted to the first unit.
in the 0-60 ppm range does not have the same effect The reactor has a diameter of 170 cm and a length of
on C2H2 and C2H4 hydrogenation. The experimental 180 cm, with a catalyst mass of 4290 kg. A typical feed
conditions were certainly different from those of the mixture consists of 4.30 wt % H2, 32.94 wt % C2H6, 51.96
present study, and this situation precludes a better wt % C2H4, 0.48 wt % C2H2, 5.33 wt % CH4, 0.08 wt %
comparison of results. Most of the kinetic studies C3H4, 1.37 wt % C3H6, 1.67 wt % C4H6, and 0.04 wt %
published to date are related to tail-end hydrogenation CO as main components. The flow rate was about
processes where low concentrations of CO and H2 are 57 900 kg/h, the total pressure was 35 atm, and the inlet
used. temperature was 342.5 K. The pressure drop was quite
Finally we need to mention that the absence of pore negligible. These conditions allow us to simulate the
diffusion limitations in the laboratory experiments was reactor with a plug flow, pseudohomogeneous, isobaric,
verified using the Weisz-Prater criterion: and adiabatic model. The high flow velocity eliminates
the need for external mass transport considerations, and
the absence of internal concentration gradients was
(-rA)obsFcatdp2
Φ) <1 (26) already demonstrated using the laboratory reactor data.
DeACsA Consequently the industrial reactor can be described by
eq 6 and a simplified form of eq 7:
where the observable rate (-rA)obs was calculated from
eq 24 at different conversion levels. Only at χA > 90% d(vCi)
does the Φ module become greater than 1, due mainly ) (-ri)Fcat (6)
dz
to the low value of CsA. For C2H4 hydrogenation the
criterion is always satisfied because -rE < -rA and the dT
conversion is less than 10-15%.
Comparison of Reactor Model III with Experi-
cp
dz
) Fcat ∑i (-∆Hi)(-ri) (7′)

mental Data. Computer simulations of the laboratory


reactor, using model III and the kinetic equations of The simulation results are presented in Figures 5 and
Table 8, are presented in Figures 3 and 4 where the 6. The predicted conversion of C2H2 is close to 100%,
conversion and the axial temperature are plotted as a but the measured value in the plant reactor was around
function of the bed length. The predicted C2H2 and 77%. In addition the actual C2H4 loss was lower than
C2H4 conversions at the reactor end are in good agree- expected. On the other hand, an exit temperature of
ment with the experimental results. On the other hand, 364.5 K was calculated, which is close to the measured
the temperature profiles show the presence of a hot spot value of 363.2 K. The pronounced disagreement be-
+ +

1504 Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996

exceeds that of C2H4 hydrogenation up to a high level


of conversion and (ii) that the presence of CO is a
necessary condition to avoid a runaway situation but
the selectivity is not dependent on the CO concentration.
Computer simulation of the laboratory reactor using
a nonisothermal model that accounts for interfacial
gradients provides a good correlation of experimental
data and indicates the existence of a temperature
overshoot near the reactor entrance. We were not able
to reproduce the conversion profile of an industrial
adiabatic converter due to the presence of a large
amount of 1,3-C4H6 in the cracked gas mixture.
Figure 5. Simulation of an industrial reactor. Acetylene and
ethylene conversion axial profiles for a model II adiabatic reactor.
9, C2H2 conversion; O, C2H4 conversion.
Nomenclature
A ) heat transfer area per unit length of the reactor, m2
Ai ) preexponential factor in Arrhenius expression,
moln-1/(m3(n-1) s)
am ) specific interfacial surface area
Ci ) concentration of ith component in the gas phase, mol/
m3
C*i ) dimensionless concentration, Csi/Ci
Csi ) surface concentration of component i, mol/m3
cp ) average heat capacity, J/(mol K)
De ) effective diffusivity
Dio ) Damkohler number
dp ) pellet diameter, m
dt ) inside diameter of tube, m
dto ) outside diameter of tube, m
Ei ) activation energy of the component i, J/mol
Figure 6. Simulation of an industrial reactor. Computed and Fi ) molar flow rate of component i, mol/s
measured temperature profiles. 9, Plant data. (-∆H)i ) heat of reaction, J/mol
h ) heat transfer coefficient, J/(m2 K s)
tween the estimated and measured conversion values hw ) wall heat transfer coefficient, J/(m2 K s)
and the approximate temperature profile is due to the hwf ) wall/fluid heat transfer coefficient, J/(m2 K s)
large concentration of 1,3-C4H6 in the feed mixture. hfs ) fluid/solid heat transfer coefficient, J/(m2 K s)
Industrial practice indicates that the hydrogenation of ki ) Arrhenius type rate constant of component i
1,3-C4H6 takes place simultaneously with C2H2 hydro- Ki ) equilibrium constant of component i
genation. The conversion of 1,3-C4H6 for the operating kmi ) mass transfer coefficient, m/s
conditions mentioned above was around 44%, and this kp ) pellet conductivity, J/(m K s)
reaction limits the activity for C2H2 hydrogenation. kref ) radial effective conductivity, J/(m K s)
The good agreement between the predicted and krf ) radial conductivity of the fluid, J/(m K s)
measured temperature profiles is due to compensation krs ) radial conductivity of the solid, J/(m K s)
of the heat of reaction of unconverted C2H2 by that ksi ) Arrhenius type rate constant of component i at the
produced in the hydrogenation of 1,3-C4H6. A certain surface
degree of catalyst deactivation, due to green oil forma- KSi ) equilibrium constant of component i at the surface
tion, may also have altered the results. A suitable kw ) wall conductivity, J/(m K s)
Pr ) Prandt number
equation for the rate of 1,3-C4H6 hydrogenation must
Re ) Reynolds number
be developed in order to predict the observed conversion
ri ) rate of reaction of component i, mol/(kgcat s)
values in the industrial reactor. As an alternative, we
[S] ) concentration of unoccupied sites
have developed an empirical model for butadiene and
Sc ) Schmidt number
propylene + methylacetylene hydrogenation, based on [St] ) concentration of free and occupied sities
industrial data. By including these equations in the SV ) space velocity, cm3 STP gcat-1 min-1
mathematical model of the industrial reactor and also T ) temperature in the gas phase, K
adding a deactivation parameter, a good correspondence T* ) dimensionless temperature of the fluid, TS/T
between the steady state experimental data and simu- Ts ) surface temperature, K
lation results was obtained (Schbib et al., 1994). Tw ) wall temperature
U ) overall heat transfer coefficient, J/(m2 K s)
Conclusions v ) average velocity of the fluid through the bed, m/s
w ) mass of catalyst, kg
Rate equations for C2H2 and C2H4 hydrogenation on yi ) molar fraction of component i
a commercial Pd/R-Al2O3 catalyst have been obtained z ) reactor length coordinate, m
for reaction conditions similar to those of a front-end Greek Letters
hydrogenation process. Experimental data as a function  ) void fraction of packed bed
of temperature, pressure, and CO content were obtained F ) fluid density, kg/m3
in a laboratory reactor covering a wide C2H2 conversion Fcat ) apparent catalyst density
range. To our knowledge, kinetic studies of this kind χi ) experimental conversion of component i
have not been reported in the open literature. The rate 〈χi〉 ) calculated conversion of component i
equations reflect (i) that the rate of C2H2 hydrogenation ηi ) effectiveness factor of component i
+ +

Ind. Eng. Chem. Res., Vol. 35, No. 5, 1996 1505


Subscripts Lam, W. L; Lloyd, L.; Catalyst aids Selective Hydrogenation of
A ) acetylene Acetylene. Oil Gas J. 1972, March 27, 66.
CO ) carbon monoxide Leviness, S.; Nair, V.; Weiss, A. H. Acetylene Hydrogenation
Selectivity Control on PdCu/Al2O3 Catalysts. J. Mol. Catal.
E ) ethylene 1984, 25, 131.
Et ) ethane Margitfalvi, J.; Guczi L.; Weiss, A. H. Reaction of Acetylene during
H ) hydrogen Hydrogenation on Pd Black Catalyst. J. Catal. 1981, 72, 185.
McGown, W. T.; Kemball, C.; Whan, D. A. Hydrogenation of
Literature Cited Acetylene in Excess of Ethylene on an Alumina-Supported
Palladium Catalyst at Atmospheric Pressure in a Spinning
Adúriz, H. R.; Bodnariuk, P.; Dennehy, M.; Gı́gola, C. E. Activity Basket Reactor. J. Catal. 1978, 51, 173.
and Selectivity of Pd/R Al2O3 for Ethyne Hydrogenation in a Men’shchikov, V. A.; Fal’kovich, Yu.G.; Aérov, M. E. Hydrogenation
Large Excess of Ethene and Hydrogen. Appl. Catal. 1990, 58, Kinetics of Acetylene on a Palladium Catalyst in the Presence
227. of Ethylene. Kinet. Catal. 1975, 16, 1335.
All-Ammar Asad S.; Webb, G. Hydrogenation of Acetylene over Moses, J. M.; Weiss, A. H.; Matusek, K.; Guczi, L. The effect of
Supported Metal Catalysts, Part 1. J. Chem. Soc., Faraday Catalyst Treatment on the Selective Hydrogenation of Acetylene
Trans. 1978, 74, 195. over Palladium/Alumina. J. Catal. 1984, 86, 417.
Battiston, G. C.; Dallaro, L.; Tauszik, G. R. Performance and Aging
Paloschi, J. R.; Perkins, J. D. An Implementation of Quasi
of Catalysts for the Selective Hydrogenation of Acetylene: A
Newtonian Methods for Solving Sets of Nonlinear Equations.
Micro pilot-plant study. Appl. Catal. 1982, 2, 1.
Comp. Chem. Eng. 1988, 12 (8), 767.
Bos, A. N. R.; Westerterp, K. R. Mechanism and kinetics of the
selective hydrogenation of ethyne and ethene. Chem. Eng. Park, Y. H.; Price, G. L. Deuterium Tracer Study on the Effect of
Process. 1993a, 32, 1. CO on the Selective Hydrogenation of Acetylene on Pd/Al2O3.
Bos, A. N. R.; Bootsma, E. S.; Foeth, F.; Sleyster, H. W. J.; Ind. Eng. Chem. Res. 1991, 30, 1693.
Westerterp, K. R. A kinetic study of the hydrogenation of ethyne Sarkany, A., Weiss, A. H.; Guczi, L. Structure Sensitivity of
and ethene on a commercial Pd/Al2O3 catalyst. Chem. Eng. Acetylene-Ethylene Hydrogenation over Pd Catalysts. J. Catal.
Process. 1993b, 32, 53. 1986, 98, 550.
Carberry, J. J. Chemical and Catalytic Reaction Engineering; Schbib, N. S.; Errazu, A. F.; Romagnoli, J. A.; Porras, J. A.
McGraw-Hill: New York, 1976. Dynamics and Control of an Industrial Front-End Acetylene
Crider, J. E.; Foss, A. S. Effective Wall Heat Transfer Coefficients Converter. Comput. Chem. Eng. 1994, 18, S355.
and Thermal Resistances in Mathematical Models of packed Schuit, G. A.; Van Reijen, L. L. The Structure and activity of metal-
Beds. AIChE J. 1965, 11 (6), 1012. on-silica Catalysts. Adv. Catal. 1958, 10, 242.
Derrien, M. L. In Catalytic Hydrogenation; Cerveny, Ed.; Studies Weiss, A. H.; Le Viness, S.; Nair, V.; Guczi, L.; Sarkany, A.; Schay,
in Surface Science and Catalysis, 27; Elsevier: Amsterdam, A. The Effect of Pd Dispersion in Acetylene Selective Hydro-
1986. genation. Proc. Int. Congr. Catal. Dechema 1984, 8th, 591.
Dixon, A. G.; Cresswell, D. L. Theoretical Prediction of Effective
Heat Transfer Parameters in Packed Beds. AIChE J. 1979, 25 Received for review September 27, 1995
(4), 663. Accepted January 19, 1996X
Doraiswamy, L. K.; Sharma, M. M. Heterogeneous Reactions:
Analysis, Examples and Reactor Design; John Wiley & Sons: IE950600K
New York, 1984.
Froment, G. F.; Bischoff, K. B. Chemical Reactor Analysis and
Design; Wiley & Sons, Inc.: New York, 1990.
Gva, L. Z.; Kho, K. E. Kinetics of acetylene hydrogenation on X
Abstract published in Advance ACS Abstracts, March 15,
palladium deposited on alumina. Kinet. Catal. 1988, 29 (2), 381. 1996.

You might also like