ind eng.chem.res.dec.2001

23
 1   Manuscript for th e special IECR issue CREVIII conference held in Barga, Italy on June 24-29, 2001 Catalytic Cracking of Cumene in a Riser Simulator A catalyst activity decay model S.Al-Khattaf (2) and H. de Lasa (1.*)  (1) Chemical Reactor Engineering Centre, Faculty of Engineering Science, University of Western Ontari o, London, Ontario, Canada N6A 5B 9 (2) Department of Chemical Engineering, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia. (*) Corresponding author Abstract The present study reports the catalytic cracking of cumene using a novel CREC Riser Simulator. Two sizes of zeolite crystals (0.4 and 0.9 μm) are employed in the experiments. The goal is to develop and test, under relevant reaction conditions for FCC (temperature, contact time, reactant partial pressure, catalyst /oil ratio), a new catalytic activity decay model based on “reactant conversion”. Kinetic and decay model parameters are estimated using regression analysis. Activation energies, deactivation constants, and Arrenhius pre-exponential constants are calculated with their respective confidence intervals. The lack of transport constrains of cumene molecules, while evolving inside the zeolites, is confirmed evaluating a modified Thiele modulus and effectiveness factors. Introduction Catalyst deactivation in FCC can take place due to several factors. Catalyst pellets can lose its shape or mass due to attrition and high temperatures (sintering). Catalyst poisoning can also take place given the effect of impurities contained in the feedstock. Typical contaminants are hydrocarbons containing S, N, O, Ni and V [1-3]. However, under normal FCC conditions coke is the most important factor affecting catalyst activity. As catalytic reactions progress coke deposits on the catalyst surface, covering active sites and leading to catalyst decay. In this respect, important modeling efforts have been addressed to the modeling of catalyst decay due to coke. A first approach towards modeling catalyst deactivation is based on a time-on-stream decay model (TOS). This empirical model assumes that the coking rate is independent of reactant composition, extent of conversion, and hydrocarbon space velocity [4]. In this respect, classical decay models include an exponential decay law, ϕ=exp (-αt) [5] and a power decay

Upload: chantran90

Post on 04-Apr-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 1/23

 1

 

 Manuscript for the special IECR issue

CREVIII conference held in Barga, Italy on June 24-29, 2001

Catalytic Cracking of Cumene in a Riser Simulator

A catalyst activity decay modelS.Al-Khattaf (2) and H. de Lasa (1.*) 

(1) Chemical Reactor Engineering Centre, Faculty of Engineering Science, University of Western Ontario, London, Ontario, Canada N6A 5B9 

(2) Department of Chemical Engineering, King Fahd University of Petroleum &Minerals, Dhahran 31261, Saudi Arabia.

(*) Corresponding author 

AbstractThe present study reports the catalytic cracking of cumene using a novel CREC Riser 

Simulator. Two sizes of zeolite crystals (0.4 and 0.9 μm) are employed in the experiments. The

goal is to develop and test, under relevant reaction conditions for FCC (temperature, contacttime, reactant partial pressure, catalyst /oil ratio), a new catalytic activity decay model based

on “reactant conversion”. Kinetic and decay model parameters are estimated using regression

analysis. Activation energies, deactivation constants, and Arrenhius pre-exponential constants

are calculated with their respective confidence intervals. The lack of transport constrains of 

cumene molecules, while evolving inside the zeolites, is confirmed evaluating a modified

Thiele modulus and effectiveness factors.

IntroductionCatalyst deactivation in FCC can take place due to several factors. Catalyst pellets can

lose its shape or mass due to attrition and high temperatures (sintering). Catalyst poisoning can

also take place given the effect of impurities contained in the feedstock. Typical contaminants

are hydrocarbons containing S, N, O, Ni and V [1-3]. However, under normal FCC conditions

coke is the most important factor affecting catalyst activity. As catalytic reactions progress

coke deposits on the catalyst surface, covering active sites and leading to catalyst decay. In this

respect, important modeling efforts have been addressed to the modeling of catalyst decay due

to coke.

A first approach towards modeling catalyst deactivation is based on a time-on-stream

decay model (TOS). This empirical model assumes that the coking rate is independent of 

reactant composition, extent of conversion, and hydrocarbon space velocity [4]. In this respect,

classical decay models include an exponential decay law, ϕ=exp (-αt) [5] and a power decay

Page 2: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 2/23

 2

law ϕ= t-n .  More recent contributions consider modified exponential functions, ϕ=exp(-αtn)

with n=0.4 [6] and n=0.5 [6]. Regarding TOS models, Szepe and Levenspiel [8] showed that

decay functions can be derived by assuming that the rate of activity decay is a function of the

number of active sites involved in the deactivation event. In this respect, dϕ /dt = - k d ϕn and 

dϕ /dt = - k d1 C2go ϕ were suggested[8]. Wojciechowski [9] considered the integration of dϕ /dt

= - k dϕn and this yielded ϕ = [1+ k dt(n-1)]-N.

A second approach to catalyst decay involves a catalyst activity decay function related

to the coke content. With this end, several relations were reported such as ϕ= exp (-δCc),

ϕ=1/(1+δCc), ϕ=1/(1+δCc)2,ϕ =1-δCc,ϕ =(1-δCc)

2 [10,11]. With the same objective, there are a

number of two parameter models based on coke concentration as follows: a) ϕ=1/(1+δCcn)

[12], b) ϕ=ϕR  + (1-ϕR )exp(-δCc) [13], c) ϕ= (B+1)/(B+exp(δCc) [14] d) dϕ/dCc = Fϕ + Gϕ (1-

ϕ) [15]. Corma et al.[16], studying catalytic cracking of paraffins, argued about interrelating

activity decay, coke on catalyst and reactant conversion using : a) dϕ/dt = -k md ϕm Ci, b) dCc/dt

= r 0c ϕ which is equivalent for m=2 and a first order rate of coke formation to ϕ= exp (-δCc).

In conclusion, while the mechanism of coke formation is not well understood yet and

there is still debate about suitable decay models, the present study proposes, a sound approach

for catalyst deactivation with a decay function based on “reactant conversion”. This decaymodel is tested, in the present study, in the context of the catalytic cracking of cumene in a

novel CREC Riser Simulator 

Modeling Cumene Cracking in the CREC Riser Simulator

Cumene cracking is a first order reaction [17]. Its disappearance in the Riser Simulator 

can be represented by the following species balance equation;

)1(Cφηk dt

dCW

VAinin

A

c=−  

where CA represents cumene concentration in the Riser Simulator (Kmole/m3), V the volume of 

the riser (45.10-6 m3 ), Wc the mass of catalyst (0.81.10-3 gcat), t the reaction time in seconds, η 

the effectiveness factor , φin the intrinsic deactivation function, k in the intrinsic rate constant in

m3/Kgcat.s.

Page 3: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 3/23

 3

 

Moreover, the concentration of cumene, CA can be expressed by the following

relationship;

)2(VMW

WyCA

hcAA =  

where yA represents the cumene mass fraction, Whc the total mass of hydrocarbons inside the

riser (0.16.10-3 Kg), MWA the cumene molecular weight (120 Kg/Kmole).

Furthermore, k in can be expressed using an Arrhenius’ relation as k in=k 0 exp(-ER /RT).

Then, eq (1) can be written as follows;

)3(yφ)

RT

Eexp(k η

dt

dy

W

VAin

R o

A

c

−=− 

In addition, and in order to reduce parameter cross-correlation it is frequent to adopt the

the following form of eq(3):

(4)Aino

R o

A

cyφ)]

T

1

T

1(

Eexp[ηk'

dt

dy

W

V−

−=−

 

where To represents the average temperature used in the reaction experiments.

It can be observed that eqs (1) and (4) account for the possible diffusion transport

limitations on the cracking reaction. In this respect, an effectiveness factor (η) has to be used to

describe the effect of diffusion on the cracking reaction [17]. Moreover and to rigorously

model this problem an effectiveness factor for an FCC catalyst can be evaluated as follows

[18]:

)5('h

)'htanh(η =

 

where h’ is the modified Thiele modulus, which is defined as;

)6(2D

Cφρ1)k (n

a

1h'

eff 

1n

Aincin

ext

+=

 

Page 4: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 4/23

 4

where ρc represents the  catalyst density in Kg/m3, n the reaction order set at one for cumene

cracking, k in the intrinsic kinetic rate constant in m3/Kgcat.sec, φin the intrinsic deactivation

function, aext the catalyst external surface area in 1/m.

However, and for the case of catalytic cracking of cumene some important

clarifications have to be introduced in eq(6) [19]. One important consideration is given by the

fact that for a first order reaction the Thiele modulus becomes independent of the reactant

concentration. Another important consideration concerns the characteristic parameter to be

considered to calculate aext. Given that essentially all the cracking reactions take place in the

zeolite crystal with little influence of the matrix, the aext should be defined using the

characteristic dimension of the crystal, L. Thus, approximating the crystal geometry with an

equivalent sphere aext = L/6 and eq(6) becomes,

)7(D

ρφk 

6

Lh'

eff 

cinin= 

with, ρc = zeolite density (825 Kg/m3)

Regarding the various kinetic parameters contained in eq(7), one very relevant

 parameter is k in. This so-called, intrinsic kinetic constant, could be hypothesized to change

with temperature following an Arrhenius’ expression, k in =k 0 exp (-ER /RT), where k 0 and ER  

represent the frequency factors and the energy of activation of the intrinsic cumene cracking

respectively.

Diffusivity in zeolites can be represented by the Eyring Equation as follows:

(8))RT

Eexp(DD D

oeff −

where ED is the diffusion activation energy. ED is estimated to be around 26 KJ/mole and D0 is

approximately 3.2x10-10 m2/sec [20]. It has to be stressed that this estimation is based on

diffusivity measurements in faujasites performed with aromatic species with a molecular size

in the cumene range. For example, the value assigned to cumene falls in an intermediate range

with respect to the values measured for xylene and 1,3,5 tri-iso-methylbenzene [21].

Page 5: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 5/23

 5

One important matter while accounting for catalyst deactivation is to consider in eq(7)

a deactivation function, ϕin. One classical approach, while describing catalyst decay is to

consider catalyst decay as a function of time-on-stream:

)9()tαexp(φ −=  

where α is a constant and t is the time the catalyst is exposed to a reactant atmosphere (time – 

on-stream).

By substituting eq(9) into eq (4), setting η=1 and integrating the resulting equation, the

following is obtained:

)10())]texp(1(V

))T

1

T

1(

Eexp('k W

exp[y o

R oc

A α−−α

−−

−=  

In addition, it is demonstrated in the present study, that a catalyst activity decay

function can be conveniently expressed as a function of converted reactant. First, coke

formation can be postulated to take place as proportion of hydrocarbon conversion. This

approach is a sound one if one considers that coke is “mainly” the result of the primary

cracking reaction steps. This means that eq(3) can be modified to express the rate of coke

formation using a stoichiometric number, νc and the corresponding molecular weight ratio as

follows:

(11))MW

MW

V

Wyφk ( ν

dt

dC

A

ccAc

c=

 

where k is the reaction rate constant of cumene cracking and Cc the coke mass fraction based

on the catalyst weight.

Eq(11) can be divided by eq (3) and it results:

)12(MW

MW

W

W ν

dy

dC

A

c

c

hcc

A

c −=

 

Since the right side of eq(12) is assumed as constant, then eq (12) can be written as,

Page 6: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 6/23

 6

(13)Ady

dC

A

c = 

where A lumps a group of constant parameters.

By integrating equation (13) between 0 and CC and between 1 and yA, the following

equation is obtained:

(14))yA(1C Ac −=  

with

)15(MWW

MWW νA

Ac

chcc−=

 

Once this first step of relating the conversion of reactant to the coke fraction is

established, the following step in the analysis is to consider a catalyst activity decay function

 based on coke concentration:

(16))δCexp(φ c−=  

However, given the relationship between the coke concentration on catalyst, Cc and the

weight fraction of cumene, yA as given by eq(14) the following equation is obtained:

(17)))yλ (1exp(φ A−−=  

where Aδλ =  

This result can be substituted in eq(4). Then, setting η=1, the following equation,

describing the rate of reactant consumption is obtained;

)18(])yyλ [1])exp(T

1

T

1[

Eexp(k'

dt

dy

W

VAA

o

R o

A

c

−−−−

=−  

Thus, the catalyst activity decay model, based on reactant conversion, given by eq (18)

involves three parameters k’0, EA, and λ. These three parameters can be defined using a

number of experiments in the Riser Simulator as it will discussed in the upcoming

experimental section.

Page 7: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 7/23

 7

Experimental Procedures

Cumene was employed, in the present study, as model reactant compound. Regarding

the FCC catalysts, commercially available Y-zeolites, small and large crystals, manufactured

 by the Tosoh company were used. Properties of these zeolites are reported in Table 1.

These Na-zeolites were ion exchanged with NH4 NO3 to replace the sodium cation with

 NH4+. Following this, NH3 was removed and the H form of the zeolites was spray-dried using

kaolin as the filler and silica sol as the binder. The resulting 60-μm catalyst particles had the

following composition: 30 wt% zeolite, 50 wt% kaoline, and 20 wt% silica sol. The process of 

sodium removal was repeated for the pelletized catalyst. Following this the catalyst was

calcined during 2 hr at 600°C. Finally, the fluidizable catalyst particles (60-μm average size)were treated with 100% steam at 760°C for 5 hr. Table 2 reports the catalyst main properties

following catalyst pretreatment. The unit cell size was determined by X-ray diffraction

following ASTM D-3942-80. Surface area was measured using the BET method. It can be

observed that the initial zeolite unit cell size of 24.5 Å was reduced, following steaming, for 

 both zeolites, to around 24.28Å.

The catalytic activity of the Y-zeolite catalyst, prepared using the various techniques

described above, was measured in a Riser Simulator using cumene as the hydrocarbonfeedstocks. The Riser Simulator is a novel bench scale with internal recycle unit. This reactor,

invented by de Lasa [22] overcomes the technical problems of the standard micro-activity test

(MAT). The catalytic conversion of cumene is modeled in the present study. The Riser 

Simulator consists of two outer shells, lower section and upper section that permits to load or 

to unload the catalyst easily. This reactor was designed in such way that an annular space is

created between the outer portion of the basket and the inner part of the reactor shell. A

metallic gasket seals the two chambers, an impeller located in the upper section. A packing

gland assembly and a cooling jacket surrounds the shaft supports the impeller. Upon rotation of 

the shaft, gas is forced outward from the center of the impeller towards the walls. This creates

a lower pressure in the center region of the impeller thus inducing flow of gas upward through

the catalyst chamber from the bottom of the reactor annular region where the pressure is

slightly higher. The impeller provides a fluidized bed of catalyst particles as well as intense gas

mixing inside the reactor.

Page 8: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 8/23

 8

The Riser Simulator operates in conjunction with a series of sampling valves that allow,

following a predetermined sequence, to inject hydrocarbons and withdraw products in short

 periods of time. Reaction products are measured by a Hewlett Packard 5890A GC with an FID

detector and a capillary column HP-1, 25 m cross linked methyl silicone with an outer diameter of 0.22 mm and an internal diameter of 0.33 microns. Coke content is evaluated, with a +/- 4%

accuracy, in separate runs. Following this, a sample of coked catalysts is exposed to an

oxidizing atmosphere with the CO2 formed being measured quantitatively. A detailed

description of various Riser Simulator components, sequence of injection and sampling can be

found in Kraemer [23].

Cumene cracking has been used extensively as a test reaction to investigate the

characteristics of any newly developed cracking catalyst [16]. The main reaction pathway of 

cumene cracking involves the cleavage of the isopropyl group, producing benzene and

 propene.

Results and Discussion

In the case of the present study two catalysts, CAT-LC (catalyst with large zeolite

crystals) and CAT-SC (catalyst with small zeolite crystals), were studied under typical

cracking conditions using cumene as a model hydrocarbon reactant molecule. It was observedthat cumene conversion increased with both temperature and reaction time (Figs 1 and 2) . An

important result was the observed mild temperature effect with a relatively more significant

influence of reaction times. For instance, at 5 seconds reaction time, cumene conversion was

23%, 32%, 38.5% and 41% when temperatures were set at 400°C, 450°C, 500°C, 550°C. In

this respect, the conversion increased 9% from 400°C to 450°C, 6.5% from 450°C to 500°C

and 2.5% from 500°C to 550°C.

Once the decay model based on the time-on-stream, in the context of the Riser Simulator was established (eq.10), it was tested under a number of operating conditions as

follows; a) four different reaction times (3, 5, 7, and10 seconds), b) four different temperatures

(400, 450, 500, and 550oC), c) two catalysts prepared with two sizes of zeolite crystals (small

crystal CAT-SC and large crystal CAT-LC), d) a single catalyst/oil ratio (catalyst/oil = 5).

Page 9: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 9/23

 9

The three model parameters ko’, ER , and α of eq(10) with η=1, were determined using

non-linear regression (MATLAB package). Table 3 reports the parameters obtained and the

limited spans for the 95% confidence interval. Tables 4 and 5 (TOS model) report the very low

correlation between k o and ER , ER  and α and the moderate correlation between ko and α for 

 both CAT-LC and CAT-SC. Fig 1 (CAT-LC) and Fig 2 (CAT-SC) show the comparison of 

the experimental and model predicted conversions (eq 10). The adequacy of the model and of 

the selected parameters to fit to data is shown in Table 3 given the 0.94-0.97 regression

coefficients.

Regarding the kinetic parameters obtained, an interesting result is the consistency of the

energy of activation and the activity decay constant, 21.35 and 27.63 KJ/mole and 0.217 and

0.239 (1/s) respectively, for CAT-LC and CAT-SC. It has to be noticed that these values of E R  

and α are in the same range to those found by Kraemer [23] while cracking alkylaromatics in

the Riser Simulator and using a time-on-stream model.

Following this, a non-linear regression involving eq(18), the same experimental data

used with the TOS model and three adjustable parameters was considered using MATLAB

software. Table 6 reports parameters obtained, while Tables 7 and 8 give the give the low to

moderate cross-correlation for the parameters corresponding to the to the CAT-LC and CAT-

SC respectively. Fig 3 reports a comparison of model results and experimental points for CAT-

LC while Fig 4 shows the same type of analysis for CAT-SC. In addition, the adequacy of eq

(18) to represent the experimental data is proven given the 0.94-0.97 regression coefficients

(Table 6).

It can be observed that the activation energy for cumene cracking based on the RC

model is in the range of 52.75-61.96 KJ/mole with these results being in excellent agreement

with the 41.8-83.7 KJ/mole activation energies reported for cumene cracking by a large

number of researchers as mentioned in an extensive literature review by Corma and

Wojciechowski [16]. In addition, the activity decay constant remains, for both CAT-LC and

CAT-SC, in a close range of 5.5-6.2 (Table 6). Hence based on this model, at 50% conversion

the catalyst have lost already 60% of its activity, while at 70% conversion there is a 72%

activity loss.

Page 10: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 10/23

 10

 

Regarding the activity decay model, based on reactant conversion, one main

assumption is given by eq (15). This assumption considers A as a constant parameter, not

affected by cumene conversion;

)19()y(1

CA

A

c

−=

 

In order to check the validity of this assumption a number of independent runs were

developed and in each of them, in addition to the measurement of cumene conversion, the coke

formed was determined. On this basis the Cc/(1-yA) group was calculated for each run

independently. It was observed that as reported in Fig 5, A (for CAT-LC) is equal to 0.6±0.2

and this for the different temperatures and reaction times used in the study. Given that the

confidence limits for A are in the ±15% range, mainly the uncertainty on Cc measurements, the

A parameter can be assumed as a constant. This validates the process of numerical integration

of eq(18) adopted to calculate the ER , k’0, and λ parameters.

Finally and considering that for CAT-LC the determined λ parameter was 5.5 and

adopting A =0.6, then it results that δ =8.33. Thus, eq(16) can be also reported as

)8.33Cexp(φ c−= .

It can be also mentioned that Cc for cumene cracking has typically a value in range of 

0.1–0.2 mass fraction. Then at Cc =0.15, ϕ = 0.2866 or 71% of the catalyst activity is lost at

0.15 % coke concentration.

Eq(1) includes an effectiveness factor . Evaluation of kinetic parameters assumes in

 both cases η=1. This important assumption has to be validated. With this end, the modified

Thiele modulus (h’) for cumene cracking was calculated using eq (7). For instance, the α 

constant, the k 0 and the ER  parameters, based on time-on-stream, were set at 0.217 (1/s), 0.0076

(m3/kgcat.s) and 23.02 KJ/mole respectively (CAT-LC in Table 3). Substituting these

 parameter values in eq(7) for a reaction temperature of 400oC and reaction time of 5 seconds,

the modified Thiele modulus was calculated as 0.095 and the effectiveness factor as 0.996.

Similarly, the calculated λ, k o and ER  parameters of eq(18) (CAT-LC in Table 6) were used to

Page 11: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 11/23

 11

recalculate the modified Thiele modulus and the effectiveness factor. For instance for a 0.22

cumene conversion, essentially the same h’ and η, with η very close to unit, were obtained.

It can thus, be concluded that cumene cracking in FCC catalysts is controlled by the

chemical reaction rate with diffusional transport constrains paying a negligible role. This is, by

the way, consistent with the 6.8Å estimated cumene critical diameter smaller than the 7.4Å,

faujasite zeolite openings [24].

In summary, the decay models considered in the present study and based on cumene

conversion and time-on-stream show to be adequate for predicting cumene conversions in a

rather large range of conversions, temperatures, and reaction times. However, the decay model

 based on reactant conversion should be preferred given its sounder mechanistic basis. It has

also to be stressed that the decay model based on reactant conversion was considered under 

cracking conditions free of transport constrains and in this context the derived decay function

can be considered an “intrinsic activity decay function.

Conclusions

a.  Catalytic cracking of cumene is a model compound that allows to test FCC catalyst

under conditions free o intra-crystalline diffusional constrains with the effectiveness

factor remaining very close to unity.

 b.  Experiments were developed in a novel Riser Simualtor using two sizes of faujasite

crystals (0.4 and 0.9 μm). The study was conducted under relevant reaction conditions

for FCC in terms of temperature, contact time, reactant partial pressure and catalyst /oil

ratio,

c.  A new catalytic activity decay model based on “reactant conversion” was proposed and

tested. This phenomenological based model is most valuable given it allows defining the

changes of chemical species without the specific extra requirement of measuring thecoke concentration,

d.  Kinetic and activity decay parameters were estimated for this model using regression

analysis, with parameters were established with their respective confidence intervals.

The consistency of the kinetic parameters for the two-zeolite crystals is an encouraging

result to support the validity of the proposed model.

Page 12: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 12/23

Page 13: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 13/23

 13

L zeolite crystal size (micron).

h’ modified Thiele modulus, refer to eqs(6) and (7) (-).

MWA molecular weight of cumene (Kg/Kmole).

MWc molecular weight of coke (Kg/Kmole). This molecular weight is set at 800 Kg/Kmole.

n reaction order or number of sites involved in the deactivation event (-).

 N parameter decay model . reference [9].

R gas universal constant (KJ/Kmole K).

r c rate of coke formation (Kmole coke/Kgcat.s).

r co rate of coke formation excluding catalyst decay (Kmole coke/Kgcat.s). 

r 2 regression coefficient (-)

t reaction time (s).

T reaction temperature (K).

To average temperature for the experiments of 475 C or 748 K.

V Riser Simulator volume (45.10-6 m3).

Wc catalyst mass (Kgcat).

Whc mass of hydrocarbon injected into the Riser Simulator (Kg).

yA mass fraction of cumene (Kg of cumene/Kg of hydrocarbons).

Greek Symbols

α  deactivation constant for TOS model (1/s).

δ constant related to coke formation , catalyst decay function based on reactant

conversion (RC model).

η  effectiveness factor for cumene cracking (-).

ϕ  apparent deactivation function.

ϕin intrinsic deactivation function .

ϕR  residual activity.

νc stoichiometric number for coke (Kmoles coke/Kmole of cumene converted)

ρc zeolite crystal density (Kg/m3).

 Abbreviations

CAT-LC catalyst manufactured with large zeolite crystals (refer to Table 2)

CAT-SC catalyst manufactured with small crystals (refer to Table 2)

CFL confidence limit

C/O catalyst over reactant ratio (Kg catalyst/Kg reactant fed)

Page 14: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 14/23

 14

 

Literature Cited

[1] Larocca, M., “Fast Catalytic Cracking with Nickel and Vanadium Contaminants”. Ph.D.

Thesis, The University of Western Ontario, London, Ontario, Canada (1988).

[2] Larocca, M., Ng, S., and de Lasa, H., “Cracking Catalyst Deactivation by Nickel and

Vanadium Contaminants”, Ind. Eng. Chem. Res. Vol, 29, pp2181, (1990).

[3] Farag, H. “Catalytic Cracking of Hydrocarbons With Novel Metal Traps”. PhD Thesis

Dissertation, the University of Western Ontario, London-Ontario, Canada, (1993).

[4] Yates, S.J., “Fundamentals of Fluidized Bed Chemical Processes”, Butlewood Monographs

in Chemical Engineering, London, England, 121 (1983).

[5] Weekman, V.W., “A Model of Catalytic Cracking Conversion in Fixed, Moving and Fluid-

 bed Reactors”, Ind. Eng. Chem. Proc. Des. Dev., 7, pp90-95 (1968).

[6] Babitz, S.M., Kuehne, M.A., Kung, H.H., and Miller, J.T., “Role of Lewis Acidity in the

Deactivation of USY Zeolites During 2-Methylpentane Cracking”, Ind. Eng. Chem. Res,

36, pp3027-3031(1997).

[7] Hopkins, P.D., Miller, J.T., Meyers, B.L., Ray, G.J., Roginski, R.T., Kuehne, M.A., and

Kung, H.H., “Acidity and Cracking Activity Changes During Coke Deactivation of USY

Zeolite”, App Catal,136, pp29-48(1996).

[8] Szepe, S. and Levenspiel, O., “Catalyst Deactivation”, Chemical Reaction Engineering,

Proceeding of the Fourth European Symp., Brussels, Pergamon Press, Oxford, 265 (1971).

[9] Wojciechowski, B.W., "The kinetic foundation and practical application of the time on

stream theory of catalyst decay", Cat.Rev.Sci.Eng., 9(1), 79-113, (1974).

[10] Froment, G. F., and Bischoff, K. B. “Kinetics Data and Product Distribution from Fixed

 bed Catalytic Reactors Subject to Catalyst Fouling”. Chem. Eng. Sci., 17, 105 (1962).

[11] Froment, G. F., and Bischoff, K. B. “Chemical Reactor Analysis and Design”, Second

Edition, J.Wiley, (1979).

[12] Gross, B., Nace, D.M., and Voltz, S.E., “Application of a Kinetic Model for Comparision

of Catalytic Cracking in a Fixed Bed Microreactor and a Fluidized Dense Bed”. Ind. Eng.

Chem. Proc. Des. Dev., 13, pp199-203 (1974).

[13] Acharya, D.R., Huges, R., and Li, K., “ Deactivation of Silica-Alumina Catalyst During

the Cumene Cracking Reaction”, Applied Catalysis A, 52, pp 115-129, (1989).

Page 15: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 15/23

Page 16: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 16/23

 16

 

Table 1

Properties of the small and large zeolites of this study

Table 2.

Properties of CAT-SC and CAT-LC catalysts used in this study

CAT-SC(Catalyst withsmall crystals)

CAT-LC(Catalyst withlarge crystals)

Unit cell size (Å) 24.28 24.28BET Surface Area (m2/g) 148 155

 Na2O (wt%) Negligible NegligibleCrystal Size (μm) 0.4 0.9

Table 3

Kinetic constants for both CAT-LC and CAT-SC based on time on stream (TOS)

Catalyst k 0.103 

(m3/Kgcat.s)

95%

CFL

ER 

(KJ/mole)

95%

CFL

α 

(1/sec)

95%

CFL

r 2 

CAT-LC 7.61 1.3 21.35 0.96 0.217 0.066 0.94

CAT-SC 7.24 1.16 27.63 0.89 0.239 0.064 0.96

 

Table 4 Correlation matrix for CAT-LC (TOS-model)

k o ER   α 

k o 1 -0.095 0.96

ER  0.095 1 -0.043

α  0.96 -0.043 1

Small zeolite Large zeolite Na2O (wt%) 4.1 0.25SiO2/Al2O3 (mol/mol) 5.6 5.7Unit cell size (Å) 24.49 24.51Crystal size (μm) 0.4 0.9

Page 17: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 17/23

 17

 

Table 5 Correlation matrix for CAT-SC (TOS-model)

k o ER   α 

k o 1 -0.125 0.96

ER  -0.125 1 -0.046

α  0.96 -0.046 1

Table 6 Kinetic constants for both CAT-SC and CAT-LC based on reactant conversion (RC)

Catalyst k 0.103 

(m3/Kgcat.s)

95%

(CFL)

ER 

(KJ/mole)

95%

CFL

λ  95%

CFL

r 2 

CAT-LC 13.0 6.3 52.75 2.88 5.5 1.9 0.94

CAT-SC 14.12 6.62 61.96 3.2 6.2 1.9 0.97

 

Table 7 Correlation matrix for CAT-LC (RC-model)

k o ER   λ k o 1 0.73 0.98

ER  0.73 1 -0.75

λ  -0.98 -0.75 1

Table 8 Correlation matrix for CAT-SC (RC-model)

k o ER   λ 

k o 1 0.82 0.98

ER  0.82 1 -0.84

λ  0.98 -0.84 1

Page 18: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 18/23

 18

Figure Captions

Fig 1. Comparison of experimental and of modeling cumene conversions for CAT-LC based

on time-on-stream (eq(10)). C/O=5. Reported experimental points are average value of at least3 measurements. Typical errors are ± 2%.

( H ) 400°C, (Δ) 450°C,. (C) 500°C, () 550° C.

Fig 2. Comparison of experimental and of modeling cumene conversions for CAT-SC based on

time-on-stream (eq(10)). C/O=5. Reported experimental points are average value of at least 3

measurements. Typical errors are ± 2%.

( H ) 400°C, (Δ) 450°C,. (C) 500°C, () 550° C.

Fig 3. Comparison of experimental and of modeling cumene conversions for CAT-LC based

on “reactant converted” (eq(18)). C/O=5. Reported experimental points are average value of at

least 3 measurements. Typical errors are ± 2%.

( H ) 400°C, (Δ) 450°C,. (C) 500°C, () 550° C.

Fig 4. Comparison of experimental and of modeling cumene conversions for CAT-SC based on

“reactant converted” (eq(18)). C/O=5. Reported experimental points are average value of at

least 3 measurements. Typical errors are ± 2%.

( H ) 400°C, (Δ) 450°C,. (C) 500°C, () 550° C.

Fig. 5. Coke selectivity for cumene conversion as a function of reaction time. C/O=5, (c)

400°C, () 500°C, (U) 550°C.

Page 19: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 19/23

 19

 

0

0.1

0.2

0.3

0.4

0.5

0.6

0 5 10 15 20

Reaction time (sec)

   C  u  m  e  n  e  c

  o  n  v  e  r  s   i  o  n

 

Fig. 1

Page 20: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 20/23

 20

0

0.1

0.2

0.3

0.4

0.5

0.6

0 5 10 15 20

Reaction time (sec)

   C  u  m

  e  n  e  c  o  n  v  e  r  s   i  o  n

 

Fig. 2

Page 21: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 21/23

 21

 

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 5 10 15 20

Reaction time (sec)

   C  u  m  e  n  e  c  o  n  v

  e  r  s   i  o  n

 

Fig.3

Page 22: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 22/23

 22

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 5 10 15 20

Reaction time (sec)

   C  u  m

  e  n  e  c  o  n  v  e  r  s   i  o  n

 

Fig.4

Page 23: Ind Eng.chem.Res.dec.2001

7/29/2019 Ind Eng.chem.Res.dec.2001

http://slidepdf.com/reader/full/ind-engchemresdec2001 23/23

 

0

0.2

0.4

0.6

0.8

1

0 2 4 6 8 10 12

Reaction time (sec)

   C  c   /   (   1  -

  y   A   )

 

Fig. 5