Copper Catalysts in Hydrogenating Nitro-benzene to Aniline

At least, Halcon International has banked its research efforts in this direction.Halcon has developed a phenol-toaniline process which the firm says h...
0 downloads 0 Views 472KB Size
I

KEKl H. GHARDAl and C. M. SLIEPCEVICH2 Department of Chemical and Metallurgical Engineering, University of Michigan, Ann Arbor, Mich.

Copper Catalysts in

Hydrogenating Nitrobenzene to Aniline Up to 400" F., the catalysts are stable and produce nearly theoretical yields of aniline. Industrial application is enhanced by availability of raw materials from petrochemical operations possibilities for controlling catalyst poisons in the feed ease of maintaining the required temperature range in fluidized bed reactors M A i v u F A C T u R m G aniline by catalytically hydrogenating nitrobenzene is attractive as raw materials are available, yields are high, and the process incurs no by-product disposal problem. Nitrobenzene can be hydrogenated over a variety of catalysts in both liquid and vapor phases. I n this study, the vapor phase hydrogenation ovkr copper catalysts was investigated. I t was found that under optimum conditions copper catalysts are stable and produce nearly theoretical yields of aniline from nitrobenzene. However, sulfur compounds and dinitrobenzenes in the feed rapidly deactivated the catalyst at all operating temperatures. Furthermore, as the highly active catalysts are stable only in a very narrow temperature range, the high heat of reaction must be removed. A fluidized-bed reactor would be ideal for this purpose. The mole ratio of nitrobenzene to hydrogen and the particle size of the catalyst are important inasmuch as they affect the actual temperature of the catalyst particle. Catalyst deactivation is essentially irreversible-Le., catalyst activity could not be restored by oxidation followed by reduction. The rate of hydrogenation of nitrobenzene in gram moles/ hour/gram of barium-promoted copper chromite catalyst at 400' F. was correlated by the following equation :

The rate controlling step appears to be the surface reaction between molecularly

Present address, 145 E. Hill Rd., Bandra, Bombay 20, India. * Present address, University of Oklahoma, Norman, Okla.

adsorbed hydrogen and nitrobenzene dissociated on two active centers. Experimental

Equipment a n d Materials. The experimental equipment is shown. The reactor and all connecting lines were of

glass to eliminate extraneous catalytic effects. Further details have been reported by Gharda (2). Hydrogen and nitrogen, 99.9% pure with traces of oxygen and water, were obtained from commercial cylinders. The various catalysts used are designated below:

Catalysts ATo.

Composition

A

CuCrOZa

B

CuCrO2 impregnated on kieselguhr (CUO, 5%; CrzOa, 2 % P Ba-promoted CuCrOzb

C D

CuO

Preparation '/$-in. unsupported pellets l/s-in. pellets

I/&.

No. Composition E CuO F

CuO

G

CuO

Preparation Supported on silica gel prepared by coprecipitation Supported on silica gel prepared by coprecipitation Impregnated on activated &toad

+ Crz03

pellets

Impregnated on silica gelC

*

Harshaw Chemical Co. Girdler Co. Davison Chemical Co. Alcoa F-10. AIR LEAK

NITROBENZENE CAPILLARY FLOYMYTTER

t; I

TO VACUUM PUMP

GAS CAP1L L A R Y FLOWMETER

TO

VENT

Reactor and connecting lines were of glass, as a steel reactor introduced extraneous catalytic effects and the studies were made at atmospheric pressure VOL. 52, NO. 5

MAY 1960

417

7n 6 -

I

GRADEL-

80

B A

GRAD

-C

E

60 L

z

70

0 w

2

a W

>

50

z

0

,1

52 w E

L \\

I

CATALYST:

G ADE

2GMS. C 20-30 MESH

60

0 0

I

0

w

>

z c

c z W u

E

0 E a W

a W

a

-1 0

5

TEMPERATURE

400.F.

8

CATALYST: 2 OMS.- C ; 20-30 MESH RATE: 1/4 OM MOLE; GRADE 8

40

FEED

NITROBENZENE

I

20

1.0

0

2.0

3.0 TIME

5.0

4.0

6.0

Four grades of nitrobenzene were used as designated below: Nitrobenzenes NO.

Preparation

A B

As-received analytical reagent gradea Prepared from reagent grade benzene, "OB, and HnSOa; free of dinitrobenzene Prepared from petroleum-derived and benzene and technical " 0 3 H1S04; about 0.1% dinitrobenzene Nitrobenzene B f 1% (wt.) pure mdinitrobenzene

D a

0

1.0

2.0

(HOURS)

Figure 1 . Nitrobenzene prepared in the laboratory from sulfur-free benzene gave maximum conversions and least catalyst deactivation

C

PER

30

Fisher Scientific Co.

Procedure. The reactor was first packed with catalyst. To maintain isothermal conditions to within 10" F., the catalyst was diluted with either 'Is-inch glass balls or Ottawa sand. After the catalyst was charged to the reactor, the system was purged with hydrogen for 15 minutes. The reactor heater was then adjusted to the desired pretreatment temperature. Generally, the reaction temperature was less than t!ie pretreatment temperature ; therefore, following pretreatment the reactor heater had to be adjusted for the reaction

3.0

k0

TIME

(HOURS)

6.0

5.0

Figure 2. The higher rate of deactivation at low hydrogen to nitrobenzene feed ratios was caused by localized temperatures and diffusion effects

temperature, and the nitrobenzene feed was then introduced. T h e products were sampled periodically and analyzed for aniline by titration in glacial acetic acid against a standard solution of perchloric acid in glacial acetic acid, using methyl violet as a n indicator. In all experiments where the catalyst stability was good the condensed reaction products contained only aniline, water, and unreacted nitrobenzene. No intermediate reduction products such as azoxybenzene or azobenzene could be detected. Results

The experimental work involved finding first factors affecting catalyst stability and activity before determining the effect of space velocity on degree of conversion. Blank runs a t GOO0 F. using only the catalyst diluents, glass balls, and Ottawa sand proved that no homogeneous reaction was taking place. Factors Affecting Catalyst Stability a n d Activity. Optimum conditions for catalyst stability were affected by particle size of the catalyst, time and tem-

perature of pretreatment with hydrogen, reaction temperature, quality of nitrobenzene, and ratio of nitrobenzene to hydrogen in the feed. Under optimum conditions, activity of the catalyst was also reproducible with great precision. Grade A nitrobenzene, even when purified by washing with dilute sodium hydroxide and redistilling, showed less stability and activity than the grades prepared in the laboratory (Figure 1). Grade B showed some improvement over petroleum-derived nitrobenzene, Grade C,from which The olefins had been removed by sulfuric acid, but which contained about 0.1% dinitrobenzene. An experiment was made in which hydrogen gas was first passed over catalyst C in separate reactor before using to hydrogenate nitrobenzene. h-o improvement in catalj-st stability was noted.

80

70

g O 60

z

z

0 ;ij 80 E

w

30-50

P

hESH

Y)

+

/-

50

> 0 z

>

7

Figure 3. The larger temperature gradients in the large catalyst particles accounted for de-

a

k

C A T A L Y S T - 2 GMS- C

60

c 40

-I 0

=

20 /hr,

IO

0

except

Flow R o l e

1.0

w h e n (iloled

6f

2.0 TIME

0

I .o

2 .o

4 18

4.0 (HOURS)

3.0 TIME

INDUSTRIAL A N D ENGINEERING CHEMISTRY

5.0

6.0

Othiiwise

Hydroqon: 3 qm. m 0 l S l h r .

30

4.0

5.0

6.0

(HJURSI

Figure 4. Above 400" to 450" F., catalyst decreased rapidly

H Y D R O G E N A T I N G NITROBENZENE I

I

I

80

I

I

TO

70 80 2

BO

0

5 so > REACTOR

TEMPERATURE

FLOWRATE

l/4 OM. MOL

+ CATALYST'

3GM.MOL

=

$

400°

GRADE B NITROBENZENE

40

H,/HR.

2 OM. OF CATALYST

L

20-30 MESH

ALL CATALYSTS SAVE ONE WERE PRETREATED F O R 2 HRS. AT

TEMPERATURE

c

SHOWN IN BRACKETS

2

+ 3 OM.

-I

AT 400° OVER 20-SO M E S H CATALYST

a

30

1.0

2.0

TIME

8 .O

4.0

s.O

ao

HYDROQEN REACTED

P R E T R E A T E D AT 450°C

FOR P"

MOLE

2

I N HYDROOEN

HOURS 2.0

1.0

S.0

5 .o

4.0

s .O

( HOURS) TlYE

Figure 5. Maximum catalyst stability was attained by pretreatment at 400' to 450' F.

The mole ratio of nitrobenzene to hydrogen was varied between 1 to 6 and 1 to 18-i.e., from twice to six times the theoretical amount (Figure 2). The deactivation a t the lower ratios of hydrogen to nitrobenzene was presumably a temperature effect associated with diffusion into the pores and heat conductivity of the catalyst. Four sizes of catalyst were studied (Figure 3). The effect of particle size was concluded to be primarily the effect of localized temperature and diffusional gradients inside the pellet. Experimental runs were made a t 50' F. reaction temperature intervals between 350' and 550' F. Activity increased with temperature ; stability decreased above 400' to 450' F. (Figure 4). T h e pretreatment consisted of passing hydrogen over the catalyst for 2 hours. As shown in Figure 5, catalysts pretreated at 400' and 450' F. attained the same steady state activity; those activated a t higher temperatures had a slightly initial activity which they lost rapidly. Apparently, the stable catalyst component is an unreduced complex between copper oxide and chromium oxide. At a higher temperature, copper is apparently formed which loses its activity rapidly. Catalysts A through G (see designations above) were investigated for their activity and stability under optimum reaction conditions. The results are summarized in Figure 6. Catalysts A , C, and F showed nearly the same stability; catalyst F was nearly twice as active as A or C . Catalysts D and E had slightly lower initial activity and were less stable than catalysts A or C. Catalysts B and G had rather low initial activity and lost it rather rapidly. Besides, catalyst G gave rise to a tarry product containing inter-

Figure 6. Copper catalysts supported on silica were adversely affected by impregnation with sodium silicate

mediate reduction products. I n the case of catalysts E and F, neither initial activity nor stability was affected by the following treatments : Removal of sodium ions by ion exchange with ammonium ions, followed by ignition a t 900' F. Removal of sulfate ions either by starting with nitrates or by ion exchange of catalyst with ammonium and nitrate ions, followed by ignition at 900' F. However, initial activity and stability were adversely affected by impregnation with additional sodium silicate. CATALYST REACTIVATION. Upon the completion of a run, the nitrobenzene flow was stopped and the hydrogen was kept on for another 15 minutes. The reactor was then purged with nitrogen. The temperature of the reactor was raised to 600' F., and air was passed gradually over the catalyst so that the temperature did not exceed 650' F. After initial reactivation had subsided, air was passed for another hour a t 600' F. The catalyst was allowed to cool in the reactor in the current of air. I t was removed from the reactor and ignited with free access to air in a n electric muffle furnace at 900' F. for 2 hours. After cooling, the catalyst was returned to the reactor, pretreated with hydrogen in the standard way at 450' F., and the reaction was carried out as usual at 400' F. The activity of the regenerated catalyst was the r =

( HOURS1

tivity was restored by treatment in hydrogen for 2 hours at 450' F. The regeneration procedure was carried out on fresh catalyst pretreated in hydrogen at 450' F. but which had not been deactivated by the reaction of nitrobenzene. The regenerated catalyst showed no change in activity. This experiment proved that there was nothing fundamentally wrong in the method of reactivation. Effect of Space Velocity on Degree of Conversion. This variable was investigated for catalyst C at only one temperature, 400 F., where the catalyst was known to be stable. The other conditions and results are summarized in Figure 7. Space velocity was varied by using different amounts of catalyst at a constant feed rate to avoid a high temperature rise and at the same time minimize mass transfer effects. For each experiment, a fresh charge of catalyst was used. Correlation of the Reaction Rate Data. The raw experimental data on conversion were graphically differentiated to obtain the rate of reaction, r . The rate was then correlated against the partial pressures of nitrobenzene, fiNOz, and hydrogen, pHz. These results are summarized in Figure 8. The rate equation r = 0.475

is a simplified form of the equation

hGGX

+

(1 i~ K N O ~ P N Kazprrz O~

same as when the reaction was interrupted to start the regeneration. This behavior is in contradiction with German practice ( 3 ) . When deactivation was caused by dinitrobenzene, the ac-

+ K ~ P A+ KH,o$H,o)~

(2)

where the K's represent the adsorption equilibrium constants for nitrobenzene, hydrogen, aniline, and water. The denominator in the above equation represents the resistance term. Because VOL. 52, NO. 5

M A Y 1P60

419

the data could be fitted by the simplified form, the resistance was essentially constant over the range. Equation 2 holds when the rate controlling step is the surface reaction between molecularly adsorbed hydrogen and nitrobenzene dissociated on two active centers. From purely theoretical considerations, such dissociation is difficult to visualize.

0 8-

? 0.6

l 04

Grade B nitrobenrens a n d hydrogen i n

/

O:

I5

D

I I

/

I

ff 10-

J

I

/



1

E q w m n o f t h e ihne

hour*

0

420

I 0 20

Discussion

As was expected, sulfur compounds in the feed deactivated the catalyst at all operating temperatures. The sulfur compounds, probably thiophene derivatives, could not be removed from the nitrobenzene; it was necessary to make nitrobenzene from a sulfur-free benzene. The deactivation of the catalyst by dinitrobenzene was unexpected; it is not of great technical significance because the nitrobenzene as usually prepared does not contain enough dinitrobenzene to affect catalyst stability appreciably. The deactivation was apparently due to strong adsorption of the product phenylenediamines, as the activity was restored by treatment with hydrogen at 450° F. Next to impurities in the feed, the most important variable affecting catalyst stability appeared to be the actual catalyst particle temperature. The mole ratio of nitrobenzene to hydrogen affected both the rate of reaction and the heat transfer characteristics of the bed. At lower mole ratios of nitrobenzene to hydrogen, not only was the amount ot heat release per unit weight of catalyst less but the heat transfer between the flowing gases and the particles was also improved. Hence, stable operation should be realized at a higher bed temperature. Similarly, at a higher mole ratio of nitrobenzene to hydrogen, catalyst stability should be improved by operating at a lower bed temperature. Both of these predictions were verified experimentally. The easier deactivation of larger catalyst particles could be explained by the larger temperature gradients present between the reaction site and the bulk of the catalyst in the larger particles. The stability of large catalyst particles was improved by operating at lower bed temperatures. Even when sulfur-containing impurities were absent catalyst deactivation was irreversible-i.e., catalyst activity could not be restored by oxidation followed by reduction. This behavior led to the conclusion that catalyst deactivation was not due to deposition of tars but was caused by slow, irreversible, thermal transformation of the catalyst structure. In this connection, it might be noted that copper chromite loses its activity in liquid phase hydrogenations if used at too high a temperature be-

Y.

025

/l

20

40

a t 450.

60

BO

r.0475

0 os

F 100

PHI

-

I

I20

WIF

Figure 7. Adverse effects of high temperature rise and mass transfer resistance were overcome by operating at sufficiently high space velocities W, weight of catalyst, grams F, feed rate, gram moles/haur

cause of reduction to metallic copper. Hence, it is believed that the stable active catalyst is essentially unreduced copper chromite and that metallic copper rapidly loses its activity. This hypothesis was verified when it was found that catalysts pretreated at temperatures over 450” F. had both a lower initial activity and a lower stability. The only extensive published work on the catalytic hydrogenation of nitrobenzene using copper catalysts is that of Brown and Henke ( 7 ) . They used copper oxide prepared either by ignition of the nitrate or by precipitation from the nitrate by hot dilute sodium hydroxide. The catalyst prepared by the latter method was superior to that prepared by the former. Their conclusions are not directly comparable with the present study. Their catalysts were not promoted or supported and hence were almost certainly reduced to metallic copper. Their catalysts were also very finely divided and yet of low activity-Le., the amount of nitrobenzene reacted per unit weight of catalyst was low. Moreover, they were concerned only with complete reaction, and hence their catalyst deactivation studies were not sensitive enough to detect small changes in catalyst activity. Furthermore. as they used a simple heated tube and measured only the exit temperature of the gases, no great reliance can be placed on their temperature values. The Germans have operated a process for the reduction of nitrobenzene to aniline over a copper catalyst ( 3 ) . Their catalyst was made by impregnating lumps of pumice stone with a slurry of basic copper carbonate in sodium silicate solution. Their operating temperatures ranged from 200’ C. at the entrance of the reactor to 350’ C. at the exit; they also used 60 moles of hydrogen per mole of nitrobenzene feed-i.e., 20 times the theoretical requirement. They revivified their catalyst by controlled oxidation with air followed by reduction with

INDUSTRIAL AND ENGINEERING CHEMISTRY

Figure 8. Agreement between data and simplified rate equation indicates that resistance to reaction remained constant over the conditions investigated p , partial pressure, atm. r, reaction rate, gram moles/hour/gram

hydrogen. However, the present authors have found that initial activity and stability of copper catalysts supported on silica were adversely affected by impregnation with sodium silicate. The only published work on the kinetics of the reaction in the vapor phase is a brief mention by Wilson in conpection with the design of a reactor ( 4 ) . The rate of the reaction is given by the equation: = 5.79 X 104 CO.678e-2.958 where r = nitrobenzene reacting, gram rules/cc. hour expressed in terms of void volume in the reactor C = concentration of nitrobenzene, gram moles/cc.

T=”K. No mention was made about the type of catalyst used; very high ratios of hydrogen to nitrobenzene were used. At these high ratios the partial pressure of hydrogen is essentially constant, and so the equation is similar to the one proposed by the present authors. Acknowledgment

A portion of this work was supported by grants from the Michigan Gas Association and Monsanto Chemical Co. Literature Cited (1) Brown, 0. W., Henke, C. O., J . Phys. Chem. 26,161,715 (1922). (2) Gharda, K. H., Ph.D. thesis, University of Michinan, Ann Arbor, Mich., v

.

1958. (3) Groggins, P. H., “Unit Processes in Organic Synthesis,” 4th ed., McGrawHill, New York, 1952. (4) Wilson, K. B., Trans. Inst. Chem. Engrs. (London) 24, 77 (1946). RECEIVED for review May 4, 1959 ACCEPTEDJanuary 11, 1960 Part of the Ipatieff Award Presentation, Division of Industrial and Engineering Chemistry, 135th Meeting, ACS, Boston, Mass., April 1959.