+ All Categories
Home > Documents > Paper - Thermal Cracking of Propane

Paper - Thermal Cracking of Propane

Date post: 08-Apr-2015
Category:
Upload: shampex
View: 916 times
Download: 3 times
Share this document with a friend
13
THERMAL CRACKING OF PROPANE Kinetics and Product Distributions ALFONS G. BUEKENS AND G ILBERT F. FROMENT Rijksuniuersiteit Gent, J. Plateaustraat 22, Gent, Belgium The thermal cracking of propane was studied in a flow apparatus between 625' and 850' C. and at atmospheric pressure. Fairly complete product distributions, including those for C4, C5, and Ca hydro- carbons, were established and a reaction scheme was deduced. The order of the propane decomposition was determiined by comparing experiments with different degrees of feed dilution and found to vary with conversion and temperature. When the rate was fitted by means of a first-order kinetic expression, the rate coefficient decreased with increasing conversion. This so-called "inhibition" was expressed mathe- matically by considering the rate coefficient to be a hyperbolic function of conversion. The activation energy of the first-order rate coefficient increases with conversion. Finally, rate equations based on the radical nature of the process a r e discussed. HE main products of the thermal cracking of propane are T ethylene and methane on one hand, propylene and hydro- gen on the other. These products may be considered to be formed by two parallel decomposition reactions of propane. Ethane, butenes, butadiene, and aromatics are also formed. All these products are most important building blocks for the petrochemical industry, so that steam cracking of propane (or light naphtha) has become the major key process of this in- dustry. The plant capacities are growing steadily. Plants producing 250,000 tons of ethylene per year are now in oper- ation and bigger units are under way. The necessity of in- creasing the ratio of ethylene to propylene in the cracker effluent has led to increased severity in cracking conditions. The design of such units requires precise kinetic equations and detailed product distributions. The literature on propane cracking is rather extensive, but the majority of the published studies present a physicochemical character, concerned with the nature of the elementary radical steps from which the over-all process is built up. They are generally carried out under static conditions, at temperatures below 650' C. and at reduced pressure. Partial product distributions were published by Frey and collaborators (1928, 1933) and Steacie and Puddington (1938). Kinney and Crowley (1954) were mainly concerned with the formation of aromatics at nearly complete propane conversion. Schutt (1 947) gives product distributions derived from industrial data. The most comprehensive product dis- tributions were published in 1931 by Schneider and Frolich (1931) with analytical techniques which bear no comparison with present-day mass spectrographic and gas chromatographic techniques. The rate of propane cracking is generally expressed as: r = k exp( -j&)Cn Some of the most representative kinetic data are summarized in Table I and represented in Figures 1 and 2. In most studies the order was found to be 1, but Martin, Dzierzynski, and Niclause (1964) and Laidler, Sagert, and Wojciechowski (1962) found an order higher than 1. Table I shows considerable spread in activation energy and frequency factor. The absence of reliable kinetic data in the temperature Table 1. Literature Data Activa- tion Energy, Temp. Frequency Gal./ Range, Authors Order Factor Mole C. Paul and Marek (1934) 1 3.98 Y 10'6 74,850 550-650 Engel et al. (1957) ... 71,000 500-590 Martin et al. (1964) 1 . ill. 3 67,000 545-600 Laidler et al. (1962) 2.58'X lOI4 67,100 530-670 ti.5 8.50 X 10'3 54,500 Steacie and Pud- dington (1938) 1 2.88 X 10la 63,300 551-602 De Boodt (19621 ... ... 53.000 700-750 Kershenbaurn ' (1967) 1 2.40 X 10" 52,100 800-1000 range encountered in industrial practice (650' to 825' C.) is striking. This work was undertaken to explain the inconsistencies and contradictions found in the literature. Experimental data were obtained in a flow apparatus for very wide ranges of temperatures and residence times, covering those encountered in industrial operation. From the product distribution a reaction scheme is derived. The over-all kinetic equation is established. Finally, radical mechanisms are discussed in an attempt to explain some of the observed facts. Apparatus The apparatus is represented schematically in Figure 3. The flow rate of propane is measured by the calibrated capillary, C, and controlled by means of the valve, Kr. In a certain number of experiments the propane was diluted with nitrogen. The flow rate of nitrogen was indicated by the flowmeter, r, and determined from the total flow rate measured by the wet-gas flowmeter and the analysis of the exit gases. The propane and nitrogen are mixed in a packed tube, D, before entering the tubular reactor, the upper part of which serves as a preheater. Upon leaving the reactor the gases are rapidly cooled, K, and analyzed in an on-line gas chromato- graph. Their flow rate is measured in a wet-gas flowmeter, G. The reactor tube, 163 cm. long, is made of chromium steel (16% Cr, no nickel) and has an internal diameter of 4 mm. Short bends permit the insertion at different heights of four VOL. 7 NO. 3 JULY 1968 435
Transcript
Page 1: Paper - Thermal Cracking of Propane

THERMAL CRACKING OF PROPANE

Kinetics and Product Distributions

A L F O N S G . B U E K E N S A N D G I L B E R T F. F R O M E N T

Rijksuniuersiteit Gent, J. Plateaustraat 22, Gent, Belgium

The thermal cracking of propane was studied in a flow apparatus between 625' and 850' C. and at atmospheric pressure. Fairly complete product distributions, including those for C4, C5, and Ca hydro- carbons, were established and a reaction scheme was deduced. The order of the propane decomposition was determiined by comparing experiments with different degrees of feed dilution and found to vary with conversion and temperature. When the rate was fitted by means of a first-order kinetic expression, the rate coefficient decreased with increasing conversion. This so-called "inhibition" was expressed mathe- matically by considering the rate coefficient to be a hyperbolic function of conversion. The activation energy of the first-order rate coefficient increases with conversion. Finally, rate equations based on the radical nature of the process are discussed.

HE main products of the thermal cracking of propane are T ethylene and methane on one hand, propylene and hydro- gen on the other. These products may be considered to be formed by two parallel decomposition reactions of propane. Ethane, butenes, butadiene, and aromatics are also formed. All these products are most important building blocks for the petrochemical industry, so that steam cracking of propane (or light naphtha) has become the major key process of this in- dustry. The plant capacities are growing steadily. Plants producing 250,000 tons of ethylene per year are now in oper- ation and bigger units are under way. The necessity of in- creasing the ratio of ethylene to propylene in the cracker effluent has led to increased severity in cracking conditions. The design of such units requires precise kinetic equations and detailed product distributions.

The literature on propane cracking is rather extensive, but the majority of the published studies present a physicochemical character, concerned with the nature of the elementary radical steps from which the over-all process is built up. They are generally carried out under static conditions, a t temperatures below 650' C. and at reduced pressure.

Partial product distributions were published by Frey and collaborators (1928, 1933) and Steacie and Puddington (1938). Kinney and Crowley (1954) were mainly concerned with the formation of aromatics a t nearly complete propane conversion. Schutt (1 947) gives product distributions derived from industrial data. The most comprehensive product dis- tributions were published in 1931 by Schneider and Frolich (1931) with analytical techniques which bear no comparison with present-day mass spectrographic and gas chromatographic techniques.

The rate of propane cracking is generally expressed as:

r = k exp( -j&)Cn

Some of the most representative kinetic data are summarized in Table I and represented in Figures 1 and 2.

I n most studies the order was found to be 1, but Martin, Dzierzynski, and Niclause (1964) and Laidler, Sagert, and Wojciechowski (1962) found an order higher than 1. Table I shows considerable spread in activation energy and frequency factor. The absence of reliable kinetic data in the temperature

Table 1. Literature Data Activa- tion

Energy, Temp. Frequency Gal./ Range,

Authors Order Factor Mole C. Paul and Marek

(1934) 1 3.98 Y 10'6 74,850 550-650 Engel et al. (1957) . . . 71,000 500-590 Martin et al. (1964) 1 . i l l . 3 67,000 545-600 Laidler et al. (1962) 2 .58 'X l O I 4 67,100 530-670

t i . 5 8.50 X 10'3 54,500 Steacie and Pud-

dington (1938) 1 2.88 X 10la 63,300 551-602 De Boodt (19621 . . . . . . 53.000 700-750 Kershenbaurn '

(1967) 1 2.40 X 10" 52,100 800-1000

range encountered in industrial practice (650' to 825' C.) is striking.

This work was undertaken to explain the inconsistencies and contradictions found in the literature. Experimental data were obtained in a flow apparatus for very wide ranges of temperatures and residence times, covering those encountered in industrial operation. From the product distribution a reaction scheme is derived. The over-all kinetic equation is established. Finally, radical mechanisms are discussed in an attempt to explain some of the observed facts.

Apparatus

The apparatus is represented schematically in Figure 3.

The flow rate of propane is measured by the calibrated capillary, C, and controlled by means of the valve, Kr. I n a certain number of experiments the propane was diluted with nitrogen. The flow rate of nitrogen was indicated by the flowmeter, r , and determined from the total flow rate measured by the wet-gas flowmeter and the analysis of the exit gases. The propane and nitrogen are mixed in a packed tube, D, before entering the tubular reactor, the upper part of which serves as a preheater. Upon leaving the reactor the gases are rapidly cooled, K, and analyzed in an on-line gas chromato- graph. Their flow rate is measured in a wet-gas flowmeter, G .

The reactor tube, 163 cm. long, is made of chromium steel (16% Cr, no nickel) and has an internal diameter of 4 mm. Short bends permit the insertion at different heights of four

VOL. 7 NO. 3 J U L Y 1 9 6 8 435

Page 2: Paper - Thermal Cracking of Propane

log Ti [E :(set)-']

0 DE BOODT . PAUL A STEACIE a LAIDLER 9 KERSCHENBAUM

BUEKENS x = 30%

x x = 60% a x = 90%

1000 900 '90 LOO

10.0 11,o lZ,O ' I 0 $ (OK)-'

4 0

Figure 1 . Arrhenius diagram

Integral valuer of rate coefficient, i

very thin thermocouples, Philips 2ABI 10, external diameter 1 mm., so that the temperature of the gas stream itself is mea- sured. The geometry is such that the value measured is close to the mean cross-sectional temperature. The furnace is cylindrical and has a length of 160 cm. and an outside diameter of 36 cm. I t consists of two hinged half cylinders, made of fireproof concrete and surrounded by a steel casing. There are two sections: a preheat and a reaction section. The preheater is a metallic cylinder insulated by a mica sheet and wound with resistance wire. Its heating capacity is 3.5 kw. The reaction section (heating capacity 6 kw.) holds 12 vertical porcelain elements upon which resistance wire is wound at a variable speed. Six of these elements have their maximum heating capacity on the top, the other six on the bottom. Both types alternate along the cylinder wall. Both series of six elements are separately connected and controlled. TO compensate for heat losses at the exit a radiant resistance wire with a heating capacity of 1.5 kw. is placed in the bottom of

Table II. Columns Used Length, Tzmp., Products

Column Material M . C. Anahred

Silica gel 2 25 Hz, 5 3 3 4

Dimethylsulfolane on Celite Carrier gas, NZ

545 (60/100 mesh) 12 40 From C1 to C4 Squalane on Chromosorb P 8 100 All HC up to

(60/80 mesh) toluene

the cylinder. Variable transformers are built in the control circuits in order to attenuate the power input during operation.

The temperatures are recorded by means of a Leeds & Northrup potentiometric recorder, Type G.

436 l & E C P R O C E S S D E S I G N A N D D E V E L O P M E N T

Page 3: Paper - Thermal Cracking of Propane

x DEBOODT T PAUL a LAIDLER o KERSCHENBAUM A STEACIE

BUEKENS 8 K O

1 : x . 30%

3:x 8 90% 2 : ~ : 60%

"C

99 1

Figure 2. Arrhenius diagram Point values of rate coefficient, k

Analysis

The reactor effluent stream was analyzed on a Wilkens Moduline 202 gas chromatograph with hot wire detector. The columns used are described in Table 11.

The carrier gas was hydrogen, unless otherwise mentioned. The peak surfaces were measured by means of a mechanical

integrator. Calibration factors determined for nitrogen, methane, ethane, ethylene, acetylene, propane, propylene, and benzene were in excellent agreement with those published by Rosie and Grob (1957).

Experimental Results

The experiments covered the following range of variables : Temperature, 625' to 850' C. Pressure, atmospheric; maximum pressure, 1.3 atm.

absolute.

Propane flow range, 0.4 to 10 moles per hour, and thus a residence time of 0.04 to 1 second and a Reynolds number of 8 to 200.

Dilution factor, 6 (molar ratio of nitrogen to propane), 0 to 10.

The analysis of the propane was as follows: propane 98.9 mole %, ethane 0.3 mole %, propylene 0.2 mole %, 2-methyl- propane 0.6 mole %, sulfur 10 p.p.m. Before a series of experi- ments the reactor surface was deactivated with CS2 at tem- peratures between 450' and 550' C.

The material balance of the experiments checked within 1%. The results are presented in two ways: in conversion us.

V/F, diagrams and in product distribution or selectivity dia- grams. The diagrams of the first type are more directed toward a kinetic analysis, those of the second type toward a study of the reaction mechanism.

V O L 7 NO. 3 J U L Y 1 9 6 8 437

Page 4: Paper - Thermal Cracking of Propane

V (

Figure 3. Flow sheet of apparatus

Figure 4. Propane conversion vs. V / F ,

438 I & E C P R O C E S S D E S I G N A N D D E V E L O P M E N T

Page 5: Paper - Thermal Cracking of Propane

' 0 ° 1 800°C

s- x

50,

25

Figure 5. Total conversion of propane and conversion to primary products as a function of V / F , at 800" C.

Figure 4 shows t.he x us. V/F, curves for temperatures ranging from 625' to 850' C. Vis the equivalent reactor volume discussed in detail below. x represents the total propane conversion. Figure 3 shows, by way of example, the conversion us. V/F, curves for propane and the principal reaction products a t 800' C.

The propylene curve goes through a maximum, resulting from the opposing effects of production and decomposition. Figures 6 and 7 show, by way of example, the product distri- butions as a function of the total propane conversion for two temperature ranges: 725' to 750' and 825' to 850' C. The temperature dependence is not very pronounced.

Product Distribution

Primary Products,. The primary products are methane, ethylene, hydrogen, and propylene, which initially are formed in nearly equal molar quantities. Ethane is also partly a primary product, but it is produced in much smaller quantities.

Table I11 gives the primary product distribution at zero conversion for several temperature ranges.

The selectivity with respect to a given product is defined below as the ratio of the number of moles of this product formed to 100 moles of propane decomposed. The hydrogen selectivity a t zero conversion was calculated from a hydrogen balance.

The classical Rice rules for radical reactions (1934) would predict at 600' C. 6057, methane + ethylene and 40% hydrogen

Table 111. Primary Product Distribution at Zero Conversion

(Moles formed per 100 moles of propane cracked) 675/ 7.251 775/ 825/

700°C. 750" C. 8GUOC. 850OC. Methane 47 47 49 48 Ethylene 48 49 50 50 Ethane 0 . 7 1.1 1 . 4 3 . 2 Hydrogen 52 52 51 53 Propylene 52 51 49 47

+ propylene; at 800' C. 63 and 37%. Experimentally the proportion is practically 50 to 50 at 800' C.

I t follows from Table I11 that there is a slight trend in ethylene and propylene selectivity with temperature. These trends may be explained by the Rice rules, according to which the formation of n-propyl radicals requires 1.2 cal. per mole more than the formation of isopropyl radicals. Ethylene and methyl radicals are the decomposition products of n-propyl radicals; propylene and hydrogen radicals originate from iso- propyl radicals. The temperature dependence of ethane selectivity is more pronounced. This may be explained by the growing importance of initiation and termination with temperature.

The initiation can be represented by

CaHs * CH3' + CzHs'

VOL. 7 NO. 3 J U L Y 1 9 6 8 439

Page 6: Paper - Thermal Cracking of Propane

725-75Q.C 9 4

+

IO

Figure 6. Selectivity diagram for temperature range 725" to 750" C.

The ethyl radicals may then form ethane. One possible termination reaction is the recombination of

two methyl radicals

2CHsO + C2He

which also produces ethane. Figures 6 and 7 show how the selectivity for ethylene and

methane increases with increasing conversion, while that for propylene and hydrogen decreases. The ethylene-propylene molar ratio rises from 1.0 a t zero conversion to 3.7 a t 85% and 6 at 95% conversion.

In the technical literature the product distribution is often referred to the amount of ethylene produced. When propane is cracked to 90% conversion under industrial conditions, 33 kg. of propylene are produced per 100 kg. of ethylene (Burke and Miller, 1965). The figure obtained in this study is 36.6. For methane the figures are 61 and 62; for the secondary products, 7 and 7 (1.3-butadiene) and 4 and 3 (butenes).

The selectivities of the principal secondary products are shown in Figure 8. Acetylene,

Secondary Products,

methylacetylene, and propadiene (or allene) are dehydrogena- tion products of ethylene and propylene. The single curve drawn for these products is to be regarded as a mean curve, since the dehydrogenations are temperature-dependent. The rate of these reactions is lowered when the reactor wall is pre- treated with CS2. At a conversion of 95% the selectivity for acetylene amounts to 1.8%-the equilibrium value-without, but only to 0.6% with pretreatment. The activation energies are estimated to be 90,000, 80,000, and 70,000 cal. per mole for acetylene, allene, and methylacetylene formation. The temperature effect on the 1-butene formation is undeniable. The selectivities go through a maximum for a propane con- version of 40 to 50%; below that conversion 1-butene is the most important secondary product. The decrease in selectivity beyond the conversion is probably due to isomerization into 2-butene and 2-methylpropene and also to dehydrogenation into 1,3-butadiene.

There are several possibilities for the formation of 1-butene. One possibility is:

2C2H4 S CH&H2CH=CHz (1)

440 I & E C P R O C E S S D E S I G N A N D D E V E L O P M E N T

Page 7: Paper - Thermal Cracking of Propane

801 825 - 850%

* '

40

30 -

20 -

IO .

.I T _c - - T T. . TT ' Z H 6 , T . ' . . ' T T T

10 20 30 40 x 5 9 . 60 70 80 90 1

Figure 7. Selectivity diagram for temperature range 825" to 850" C.

Several objections may be formulated against this reaction as the principal source of 1-butene. Indeed, the conversion to 1-butene is higher than the equilibrium conversion of Reaction 1. Furthermore, the selectivity for 1-butene increases with temperature, whereas the equilibrium conversion decreases with temperature. Finally, Reaction 1 was found by Krauze et al. (1935) to have an activation energy of 37,700 cal. per mole, considerably less than that for total propane decom- position. A comparison of selectivities at equal conversion for different temperatures indicates that the activation energy of 1-butene formation is approximately 15,000 cal. per mole higher than that of propane decomposition, which varies between 52,000 and 64,000 cal. per mole.

For all these reasons, a more plausible way for the pro- duction of 1-butene is the recombination between methyl and allyl radicals:

ki2

CH3" C3Hb0 --t C4H3 (2)

where the allyl radicals are formed according to:

C3Hrj + R" -+ C3Hb" + RH (3)

Reaction 2 requires no activation energy. Therefore, the temperature dependence of the 1-butene formation has to be

0

traced back to the temperature dependence of the radical concentrations. When compared at the same inhibition level, these concentrations vary with temperature according to an Arrhenius-type relation, so that an activation energy may be used to characterize their temperature dependence. From the simulation of the cracking of propane on a digital computer a value of about 48,000 cal. per mole may be derived for the "activation energy" of the methyl radical, while that for the allyl radical must be between 20,000 and 30,000 cal. per mole (Buekens, 1967). I t follows that the apparent activation energy for 1-butene formation according to Reaction 2 must lie between 70,000 and 80,000 cal. per mole, in agreement with the experimental results.

As mentioned above, 1-butene isomerizes partially into 2-methylpropene and 2-butene. The selectivity for 2-butene does not exceed 0.3%. 2-Methylpropene could not be separated completely from I-butene by gas chromatography. For propane conversions of 90 to 95% the amount of 2-methyl- propene was estimated to represent some 40% of the sum of 1-butene and 2-methylpropene.

The 1,3-butadiene selectivity, also represented in Figure 8, was found to be practically independent of temperature. If 1,3-butadiene were formed by dehydrogenation of 1-butene, the selectivity would decrease at high propane conversions,

VOL. 7 NO. 3 J U L Y 1 9 6 8 441

Page 8: Paper - Thermal Cracking of Propane

2.6 .

2.4 .

s 2.0 -

* 2 1.6 - t- u W -I W

t

* 1.2 -

0 .8 .

//cH~ CH

Figure 8. Selectivity diagram for secondary products

2 CH3' * CzHs whereas the opposite is observed. 1,3-butadiene is mainly formed according to:

I t is therefore likely that

2C2H4 + CHFCH-CH=CHZ + Hz

a conclusion set forward by Schneider and Frolich (1931). The Cs and Cg hydrocarbons are mainly dienes and aromatics and are probably formed by condensation reactions of ethylene and propylene. The following products were identified : pentadiene, isoprene, methylpentadienes, cyclic dienes such as cyclopentadiene and probably cyclohexadiene; olefins such as I-pentene, 1-hexene, and 4-methyl-I-pentene; cyclic olefins such as cyclopentene and cyclohexene; and aromatics such as benzene and toluene.

The formation of aromatics increases rapidly at high con- versions. I t is strongly dependent on wall effects.

Reaction Scheme. To conclude this discussion of the product distribution, the following reaction scheme may be set up :

Hz + CzHz 7

CHI + CzH4 - '/z(Hz -t GHs) +

7 \

L / C3HS 4 H z + CsHs 3

Hz + C3Hs * '/2(Hz + Ce"o)

The reactions of this scheme probably proceed over radicals An additional set of reactions is required to almost entirely.

explain the formation of ethane and olefins:

CzHs' + R H 7 - 1 CH3' + C3H5' - I-butene + 2-butene + 1,3-butadiene +

J. Hz 2-methylpropene

C2&' + C3H5' -+ I-pentene

C3H70 + C3H50 - 4-methyl-1-pentene { l-hexene

Kinetic Study

Equivalent Reactor Volume. One of the main problems encountered in the derivation of rate equations for homogene- ous gas-phase reactions from experiments in tubular flow reactors is the longitudinal temperature profile. Indeed, it is not possible to distinguish sharply between preheat and reaction sections as can be done in fixed-bed catalytic reactors. If the rate is to be determined at a reference temperature, say TR, and if the reaction volume is counted from the point where TR is reached, an error is introduced by neglecting the conversion accomplished in the section below TR. A similar situation occurs at the outlet, where the conversion continues to some extent at a temperature lower than TR. To correct for this and to compare experimental data at the same reference temperature TR, use can be made of the equivalent reactor volume concept introduced by Hougen and Watson (1947). The equivalent reactor volume, V, is defined as that volume which, at reference temperature TR, would give the same conversion as the actual reactor, with its temperature profile. I t follows that

TT

~ T R dV = - dV' (4)

442 I & E C P R O C E S S D E S I G N A N D D E V E L O P M E N T

Page 9: Paper - Thermal Cracking of Propane

and

LvLxp( - A ) d V t

exp( -") R T R V =

The usefulness of this concept was discussed and demonstrated by Froment, Pijcke, and Goethals in their study of the thermal cracking of acetone (1961). T o be practical its utilization requires a good evaluation of the activation energy, E, prior to the knowledge of the rate coefficient, k.

The experimental data of the present work were corrected by means for this concept, using an estimated value of 50,000 cal. per mole for the activation energy. This value was sufficiently close to the final value obtained from an Arrhenius diagram to make iteration superfluous. In some cases the true volume was reduced by a factor of 2.

Determination of Order a n d Rate Coefficient, In this work it was attempted, as is usual in the study of cracking reactions, to describe the rate by a simple equation like

r = kCn ( 5 )

When the differential method of analysis is applied, the rate is obtained as the slope of the tangent to the curves of x us. V/Fo. When the integral method of analysis is used, Rate Equation 5 has to be substituted in the continuity equation

F,dx = rdV

and the resulting equation integrated. The data points x us. V/F, are then used to calculate k.

I t was found using both methods of analysis that the rate coefficient decreases with increasing conversion, a t least for orders between 1 and 1.5. Using the terminology of the cracking literature the reaction is "inhibited by its reaction products or by intermediate species." Another way to put it is that Rate Equation 5 does not adequately describe the rate of the complex cracking phenomenon. Yet, in this section, the kinetic treatment is based upon Equation 5 , whose simplicity is attractive for practical purposes. Moreover, this treatment leads to considerable insight into more complex mechanisms. The variation of the rate coefficient, k, defined by Reaction 5 , with conversion necessitates a distinction between point and integral values. The differential method leads to point values; the integral method to integral values, E .

I t also follows that the correct order of the reaction can be obtained only by comparison of data having the same level of inhibition, which is mainly determined by the relative rates of reaction of the chain-carrying radicals with propane and propylene. For this reason k/k,, the ratio of the point value of the rate coefficient a t conversion x to the point value at zero conversion, may be written:

k k(R") (C3Hd k, - k(Ro)(C3Hd + k"(Ro)(C3Hd - E

Since the selectivity for propylene is almost independent of the pressure, the extent of inhibition is in first approximation determined by the conversion, so that Equation 6 is obtained:

Therefore, the determination of the order requires compari- son of experiments carried out a t different ratios of nitrogen to propane but leading to the same conversion. The procedure

Table IV. Variation of Order

c. c. C. 750" C. C. C. 625' 650' 675' 700- 775' 825'

Conversion 0.022 0.058 0.067 . . . 0.238 0.412 Order 1.36 1.21 1.14 1 0.99 0.96

used in this work is based upon the integral method of analysis. The formulas used to calculate & for orders of 0.5, 1, 1.5, and 2 are given in the Appendix. I n these equations the expansion factor is taken to be 2 over the complete range of conversion, as observed experimentally. Figure 9 illustrates how the correct order is obtained from the intersection of two curves related to the experiments being compared.

The semilogarithmic representation used leads to a sharper intersection. Table I V shows the mean values of the order a t several temperatures.

Table IVdoes not make it possible to conclude unambiguously whether the order varies with conversion, with temperature, or with both. Indeed, the conversions used to determine the order at the higher temperatures are much higher than those at the lower temperatures. I t has been shown, however, that the order decreases mainly with increasing conversion-Le., with increasing inhibition (Buekens 1967).

I n the preceding section the order of the reaction, although not constant, was found to be close to 1. With a fixed value of 1 for the order it becomes possible to determine a n activation energy and a frequency factor for propane decomposition. When the first-order & values, deter- mined by the integral method, are plotted in an Arrhenius diagram (Figure l ) , a definite trend with respect to conver- sion is observed. The following activation energies and frequency factors are obtained:

First-Order Kinetics.

30% conversion

58,000 130 = 2.33 X 10l2 exp( -T) (second-')

60% conversion

= 8.62 X lo1* exp ( -- 6yio) (second-')

90% conversion

= 1.64 X 1013exp ( -- ':io) (second-')

This shift in activation energy may explain the spread of the values reported in the literature, probably obtained at different levels of inhibition.

I t is of importance to present point values of the rate co- efficient, too. Indeed, most of the k values reported in the literature are point values. Furthermore, if a relation is to be established between the over-all rate coefficient and the inti- mate nature of the reaction, point values are of more direct interest than integral values. Equation 6 suggests a hyper- bolic law for the variation of point values k, with conversion:

k0

1 + ax k = - (7)

In Equation 7, a is an empirical inhibition coefficient and k, is the value of k a t zero conversion and therefore a rather artificial quantity. Indeed, k should vary with conversion in the same way as the radical concentration. The latter goes through a maximum: I t increases as the reaction starts, but decreases

VOL. 7 NO. 3 J U L Y 1 9 6 8 443

Page 10: Paper - Thermal Cracking of Propane

+:

+ I

0

- 1

- 2

ORDER i 1/2 1 I20 3/2

Figure 9. Determination of reaction order

t R = 675OC. V / F , = 7.86 x = 0.0253, dilution factor 6 = 0 V / F o = 13.50 x = 0.0259, dilution factor 6 = 6.43

afterward, as a result of the formation of allyl radicals. Such a maximum was observed experimentally a t the very low con- versions which were measured in the temperature range 625’ to 675’ C. The hyperbolic law of Equation 7 does not con- sider the drop in k as zero conversion is approached, of course. The “initial value,’’ k,, and the inhibition coefficient, a, the constants of the empirical hyperbolic law describing the in- hibition, were determined in the following way.

When the expansion factor equaIs 2, the combination of Equation 7 with the continuity and rate equations leads to:

(8) dx - kOCt(1 - x ) - - v

d- FO

(1 + a x ) ( l + x )

Upon integration and rearrangement, Equation 8 gives:

C,k, - = -a 2 In (1 - x ) + 2 x + - - 2 In (1 - 2) + x

(9)

V X 2

FO 2

Temg . , O c. 625 650 675 700 725 750 775 800 825

Table V. k,, a VI. Temperature

kQ, Sec. -1 0.0548 0.1819 0,3298 0.9452 1 ,7800 2.2140 4,0760 9.3580

13.9500

a

7.345 11.720 3.696 6.505 5.896 2.371 1.128 1.799 1.375

With an equation of this form k, and a may be calculated from a set of experimental x us. C,V/Fo values by linear re- gression.

Figure 10 shows a k us. V/Fo or x curve and compares experi- mental and calculated E values at 775’ C. The calculated E values were obtained from:

The results are shown in Table v.

444 l & E C PROCESS DESIGN AND DEVELOPMENT

Page 11: Paper - Thermal Cracking of Propane

I r V / F o . . v k d - k = v - J, FLl

FO

and the k values were related to V/Fo through Equations 7 and 9.

I t follows from Figure 10 that the hyperbolic law (Equation 7 ) permits an excellent description of the inhibition.

I t is seen from Table V that the values of a are less in line at the lower temperatures of 625' to 675' C. a is very sensitive to the spread on the data. The maximum in the k us. V/Fo curves also influences the results. From an Arrhenius diagram, the following temperature dependence is derived for a:

a = 3.01 X lO5exp ___ (2:3 If a were exactly equal to the ratio kN/k' , the temperature

dependence would be less pronounced and amount to no more than 3000 cal. per mole. The value of 23,100 may be ex- plained by the role played by the allyl radical, C~HS' , which becomes more effective as a chain carrier a t higher temperature.

The following temperature dependence is found, by linear regression, for k,:

k, = 4.10 X 10" exp ( - - 'y:) (second-')

I t now becomes possible to calculate point k values a t These values

The following activation energies and frequency factors are

different conversions-e.g., 30, 60, and 90%. are plotted in an Arrhenius diagram, as shown in Figure 2 .

found by linear regression:

30y0 conversion:

k30 = 5.10 X 1OI2 exp ( - ~ 'y;) (second-')

6OYc conversion :

k60 = 1.98 x loi3 exp ( - __ 6y:o) (second-')

90% conversion:

k g o = 3.84 x 1013 exp ( - ~ 6y:) (second-')

The point values are lower than the integral values, of course, while the activation energies are higher, as may be foreseen from the trend observed in Figure 1 .

The activation energies given here are lower than those usually reported. No effect of heat transfer limitations may be involved: The thermocouples are inserted in the gas stream itself, not on the wall. However, De Boodt (1962) obtained a value of 53,300 cal. per mole in a flow apparatus of different construction than the one used in this study, while Kershen- baum and Martin (1967) came to an expression for the first- order rate coefficient in the temperature range 800' to 1000' C. (Table I) which is in remarkable agreement with the equation given here for k,.

The over-all order of propane de- composition depends upon the conversion and temperature. Values ranging from 1.3 to 1 were obtained. This clearly indicates that a simple power law like Equation 5 is not ade- quate for the description of the rate of propane decomposition under widely varying conditions.

More satisfactory equations have to be based upon a more detailed consideration of the nature of the reaction.

Radical Mechanisms.

7 7 5%

Figure 10. Comparison of experimental & values with those calculated from point values

I n 1934 Rice proposed the following sequence to describe the thermal cracking of propane under initial conditions:

With this system of reactions the rate of propane decompo- sition may be written:

I.' where 0 = C1 -

FO

The inaccessible radical concentrations may be eliminated from this expression with the help of the so-called steady-state approximation. According to this concept, introduced by Bodenstein, the radical concentration does not vary rapidly beyond the initial startup period. A balance on the pro- duction and consumption of CH3' and H' then leads to ex- pressions for the concentrations of these radicals as a function of the propane concentration. After substitution of these ex- pressions into Equation 10 the following rate law, which is of first-order with respect to propane, is obtained, provided no distinction is made between n- and isopropyl radicals and k i << k3.

A shift in reaction order, as experienced in this work, has been reported by Laidler, Sagert, and Wojciechowski (1 962). According to these authors the order 1.5 they observed would

VOL. 7 NO. 3 J U L Y 1 9 6 8 445

Page 12: Paper - Thermal Cracking of Propane

be due to a bimolecular initiation and a termolecular termi- nation involving two CHaO or two H o and a third body, while order 1 would be due to bimolecular initiation and termolecu- lar termination involving CH3" and C3H7' or HO, CsH7', and a third body. There appears to be little justification for such hypotheses. Indeed, initiation has always been considered to be unimolecular. I t is also widely accepted that combin- ations of two CH3' or of CH3' and C3H7O do not require a third body. There is, however, no need to resort to such hypotheses to explain the shift in the order: I t suffices to add a second termination reaction like

to the scheme proposed by Rice to come to a rate equation which is much more complex than Equation 11 and which leads to an order varying bettween 1.5 and 1 .O. Termination according to Reaction 7 must be preponderant over that proposed by Rice, because very little butane is formed in the thermal cracking of propane.

The Rice scheme, eventually completed with termination Reaction 7 , may give an excellent description of the initial order and product distribution, but it does not account for the for- mation of secondary products, nor for the inhibition. From the discussion of the product distribution it appears that the allyl radical plays an important part in the cracking of propane. This radical may be formed by hydrogen abstraction from propylene according to:

8 C3He + H o -+ C3H5' + H P

CsH6 + CH3' 5 C3Hb0 + CH4

I n turn allyl radicals may abstract hydrogen from propane according to:

C3Hs + C3H6' 2 CsH7' + C3H6

If allyl radicals were incapable of H-abstraction, Reactions 8 and 9 would have to be considered as terminations. Ter- minations sensu stricto involving allyl radicals could be :

C3H5O + CH3O C4H8 13

2C3H5' + products

Abstractions 8 and 9 and terminations 11, 12, and 13 would

A classical steady-state treatment applied to the Rice scheme be responsible for the inhibition.

leads to the rate equation:

r = + 2ki(C3Hs)

provided CH30 is the only chain-carrying radical, Reaction 7 the only termination, and Reaction 9 the only inhibition. This equation leads to a variable order that becomes one at high conversions, when kg(C3H6) predominates. Under these conditions the activation energy would be E = El + E3 - E9 = 82,000 + 8500 - 7700 = 82,800 cal. per mole, much higher than the experimental values. This may throw some doubt upon the proposed reaction scheme, but it is much more likely that the application of the steady-state approximation is not justified in the case of an inhibited system, which is character- ized by a decrease in concentration of active radicals, as a result of their reaction with the products. I t therefore seems neces- sary to attempt a quantitative explanation of the results of this

2kik3(CsHs)'

ks(C3Ha) + ~ ~ C I ~ ( C ~ H B ) ~ X 8kikdCaHs)

work which would not rely upon the steady-state approxi- mation. This requires the integration of a set of ordinary differential equations describing the evolution of the different species in the reactor. Such a computer simulation of the thermal cracking of propane is being investigated at present in our laboratory.

Appendix. Calculation of Integral Rate Coefficient for Different Orders and E = 2.

n = 0.5.2 = 1

V - dZ FO

+ (Arc sin 2x mj + 6 - Arc sin -)] 6

2 + 6

n = 1.5. =

I 2 ( 2 + 6) 1 d"f" - 1 - dl+s 1 + I -?.

4 L ' - '. 1

1 [ d(1 - x ) ( l + 6 + x ) - d i x ] -

2 + 6

2% + 6 ( 3 + :)[Arc sin 2+6 - Arc sin ~ (

T I

1 n = 2 . E = - V 1 - %

- cz Fo 1 x + 2(2 + 6) In (1 - x )

Ac knowledgment

One of the authors (A.G.B.) is grateful to 1.W.O.N.L.- I.R.S.I.A. for a research grant over the period 1965 to 1967 and to the Robert De Keyser Foundation of the Belgian Shell Co. The experimental and computational assistance of P. Martin and R. Van Slembrouck is also acknowledged.

Nomenclature

a c ct E FO k rG ko ko

k ' k" n r TT ~ T R R (RO) = T = T E =

v = x =

inhibition coefficient, dimensionless concentration, moles/cc. total concentration, moles/ cc. activation energy, cal./mole molar feed rate of propane moles/sec. rate coefficient, sec. -1 cu. m.(n-') mole'-n rate coefficient, as determined by integral method initial rate coefficient rate coefficient for dehydrogenation of ethylene into

rate coefficient for abstraction on propane rate coefficient for abstraction on propylene order of reaction rate of reaction, kmole/cu. meter/sec. rate of reaction, a t temperature T rate of reaction at reference temperature TE gas constant, cal. mole-' O C.-l concentration of Ro radicals, mole/cu. meter temperature, O C. or O K. reference temperature, O C. or O K. conversion, dimensionless equivalent reactor volume, cu. meters

acetylene

446 l & E C P R O C E S S D E S I G N A N D D E V E L O P M E N T

Page 13: Paper - Thermal Cracking of Propane

V' e = expansion factor 6 = dilution factor 6 q = selectivity, %

= actual reactor volume, cu. meters

= C, V / F , reciprocal space velocity, sec.

literature Cited

Buekens, A. G., Ph. D. thesis, Rijksuniversiteit Gent, 1967. Burks, P. D., Miller, R., Chem. Week 18, 64 (1965). De Boodt, H., Ingeniersthesis, Rijksuniversiteit Gent, 1962. Engel, J., Combe, A,, Letort, M., Niclause, M., Compt. Rend.

244. 453 (1957). Frey, b. E., Heppj H. J., Znd. Eng. Chem. 25,441 (1933). Frey, F. E., Smith, D. F., Znd. Eng. Chem. 20, 948 (1928). Froment. G. F., Piicke, H., Goethals, G., Chem. En,<. Sci. 13, 173,

180 (1961).

Vol. 111, TViley, New York, 1947. Hougen, 0. A., TVatson, K. M., "Chemical Process Principles,"

Kershenbaum, L. S., Martin, J. J., A.Z.Ch.E. J . 13, 148 (1967). Kinney, C. R., Crowley, D. J., Znd. Eng. Chem. 46,258 (1954). Krauze, M. V., Nemtzow, M. S., Soskina, E. S., J . Gen. Chem.

Laidler, K. J., Sagert, N. H., Wojciechowski, B. TY., Proc. Roy.

Martin, R., Dzierzynski, M., Niclause, M., J . Chim. Phys. 61,

Messner, A. E., Rosie, D. M., Argabright, P. A., Anal. Chem. 31,

Paul, R. E., Marek, L. F., Znd. Eng. Chem. 26,454 (1934). Rice, F. O., Herzfeld, K. F., J . A m . Chem. Soc. 56, 284 (1934). Rosie, D. M., Grob, R. L., Anal. Chem. 29, 1263 (1957). Schneider, V., Frolich, P. K., Znd. Eng. Chem. 23, 1405 (1931). Schutt, H. C., Chem. Eng. Progr. 43, 103 (1947). Steacie, E. W. R., Puddington, T. E., Can. J . Res. B 16, 411

USSR 5 , 343 (1935).

Sac. A270, 242 (1962).

286 (1964).

230 (1959).

(1938).

RECEIVED for review October 19, 1967 ACCEPTED January 26, 1968

REMOVAL OF NITROGEN FROM ARGON

WITH TITANIUM-METAL SPONGE M A R T I N L. K Y L E , R . D E A N P I E R C E , L E S i T E R F . C O L E M A N , A N D J O H N D. A R N T Z E N

Chemical Engineering Diuision, Argonne National Laboratory, Argonne, I l l . 60439

The rate of reaction of titanium-metal sponge with nitrogen in argon-nitrogen gas mixtures was studied at 900" C. At least three titanium-nitrogen solid phases are formed as the reaction proceeds, and the rate-controlling mech- anism is believed to be the diffusion rate of atomic nitrogen through the TiN,(G) phase. A single relation has been developed which describes the titanium-nitrogen reaction kinetics of argon-nitrogen mixtures. Additional mathematical relationships were developed to permit estimation of the required size and useful lifetimes of titanium-sponge gettering beds designed to remove a nitrogen impurity from otherwise pure argon.

The reaction rate is dependent on the partial pressure of nitrogen in the gas phase.

HE increasing utilization of large inert-gas enclosures has Tied to interest in purification systems capable of removing oxygen, nitrogen, and water vapor contaminants from large volumes of inert gas, particularly helium and argon. Chemical gettering methods are often employed to maintain a pure atmosphere. In a system of this type, an activated sorbent such as molecular sieves (product of the Linde Division, Union Carbide Corp.) can be used to remove water vapor, and a rela- tively low-temperature gettering material (such as manganous oxide at 150" C.) can be used to remove oxygen. The removal of nitrogen presents a greater problem because of its lower chemical reactivity.

Interest in the use of titanium as a nitrogen-gettering material has stemmed from several advantages that titanium possesses over gettering materials now commonly used. Titanium-metal sponge, a high-surface-area form suitable for gettering opera- tions, is available a t a relatively low cost and requires no pre- treatment before use. Neither the oxide nor nitride reaction products of titanium are pyrophoric, and they can be handled in air. Both titanium and its reaction products have low toxicity.

To investigate the possible utilization of titanium as a get- tering material for nitrogen impurities in argon, a study was

made of the kinetics of the titanium sponge-nitrogen reaction with nitrogen present in otherwise pure argon. A reaction temperature of 900" C. was chosen for this study asacompromise between good reaction kinetics and ease of containment of the titanium sponge. Previous work (Wasilewski and Kehl, 1954) showed that the titanium-nitrogen reaction rate increases with temperature from 750" to 1400' C. Holvever. the exist- ence of a titanium-nickel liquid phase (Hansen and Anderko, 1958) above about 955°C. prevents containment of titanium sponge in stainless steels or other nickel alloys at the higher temperatures

Equipment

A diagram of the experimental system is shown in Figure 1. An argon-nitrogen mixture of carefully controlled composition was continuously circulated through a small bed of titanium sponge. This bed was composed of eight removable sections hereafter called sample beds. Each sample bed consisted of 1 to 10 grams of -8 +IO-mesh titanium sponge (see Table I ) contained in a 2-inch 0.d. by 0.5-inch high circular tray, the bottom of which was constructed of two layers of 70-mesh stainless steel (Type 310) screen. Figure 2 shows the titanium sponge in a sample bed. Lava (aluminum silicate) spacers were used between the individual sample beds to facilitate their removal.

VOL. 7 NO. 3 J U L Y 1 9 6 8 447


Recommended