model predictive control of a fluid catalytic cracking

20
* To whom correspondence should be addressed, Fax: +30 2310 498380, E-mail: [email protected] Model Predictive Control of a Fluid Catalytic Cracking Pilot Plant G.M. Bollas 1,* , S. Avramidis 1 , S.A. Papadopoulou 2 , A.A. Lappas 1 , S.S. Voutetakis 1 and P. Seferlis 1,3 1 Chemical Process Engineering Research Institute, Centre for Research and Technology Hellas, 6th km Charilaou–Thermi Rd, Thessaloniki, PO Box 361, GR-570 01, Greece; 2 Automation Department, Alexander Technological Educational Institute of Thessaloniki, PO Box 14561, GR-541 01, Thessaloniki, Greece; 3 Department of Mechanical Engineering, Aristotle University of Thessaloniki, PO Box 484, GR-54 124, Thessaloniki, Greece Abstract The main objective of this work is to improve the control performance during catalyst benchmarking experiments in the Fluid Catalytic Cracking (FCC) pilot plant operated in the Chemical Process Engineering Research Institute (CPERI). Catalyst benchmarks require unit operation within a predefined span in order to match specified operational conditions that are requested by the industry as standard. In order to diminish cases of repetitive experiments, a non-linear Model Predictive Control (MPC) strategy is implemented for the tight and efficient control of the pilot process. In an earlier work [1] a dynamic mathematical model has been developed for the simulation of the dynamic behavior of the FCC pilot plant of CPERI. The riser simulator predicts the conversion, coke yield, and heat consumed by the catalytic reactions in the pilot riser on the basis of semi-empirical models developed in CPERI [2, 3]. The pilot regenerator model uses the two-phase theory of fluidization [4] with a dilute phase model in account for post-combustion reactions [5, 6]. The dynamic simulator has been validated against steady- state and dynamic experiments performed in the pilot plant [1]. For the case of the pilot plant, the variables that may be manipulated are the catalyst circulation rate and feed preheat temperature, whereas the effects of feedstock composition and catalyst quality are considered as disturbances. The controlled variables may be the riser temperature or/and the feed conversion. These two variables differ in nature regarding their controllability aspects. The riser temperature is an online measured variable, whereas the feed conversion is a variable that can only be measured after one hour of steady state plant operation. Additionally, constraints may be imposed on the CO, NO x , and SO 2 emissions of the regenerator for their partial control. As the pilot plant regenerator operates under full combustion mode the goal of minimum CO emissions can be easily achieved, yet for the other goals, the potential of an optimal operating point of the pilot plant may be explored. Typically, the MPC strategy involves specified prediction and control horizons. In such a scheme the optimal piecewise constant future control actions are calculated through dynamic optimization over specified time intervals. Multi-rate process measurement updates are obtained frequently, whereas the unobservable controlled variables (the feed conversion in this study) can be inferentially estimated using the dynamic model and the process measurements. The sensitivity of the performance of an optimal MPC structure in respect to the number of control intervals and the length of prediction and control horizons was examined in conjunction with the computational time necessary for the numerical solution. The MPC scheme allowed for an accurate targeting of the desired feed conversion with limited knowledge about the catalyst activity and selectivity, while showing robustness during simulation. In conclusion, the MPC strategy provided tight following of the prescribed operating conditions and objectives and promises the final reduction of redundant experiments with the same catalyst in catalyst evaluation tests.

Upload: others

Post on 11-Apr-2022

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Model Predictive Control of a Fluid Catalytic Cracking

* To whom correspondence should be addressed, Fax: +30 2310 498380, E-mail: [email protected]

Model Predictive Control of a Fluid Catalytic Cracking Pilot Plant

G.M. Bollas1,*, S. Avramidis1, S.A. Papadopoulou2, A.A. Lappas1, S.S. Voutetakis1 and P. Seferlis1,3 1Chemical Process Engineering Research Institute, Centre for Research and Technology Hellas, 6th km Charilaou–Thermi Rd, Thessaloniki, PO Box 361, GR-570 01, Greece; 2Automation Department, Alexander Technological Educational Institute of Thessaloniki, PO Box 14561, GR-541 01, Thessaloniki, Greece; 3Department of Mechanical Engineering, Aristotle University of Thessaloniki, PO Box 484, GR-54 124, Thessaloniki, Greece

Abstract The main objective of this work is to improve the control performance during catalyst

benchmarking experiments in the Fluid Catalytic Cracking (FCC) pilot plant operated in the Chemical Process Engineering Research Institute (CPERI). Catalyst benchmarks require unit operation within a predefined span in order to match specified operational conditions that are requested by the industry as standard. In order to diminish cases of repetitive experiments, a non-linear Model Predictive Control (MPC) strategy is implemented for the tight and efficient control of the pilot process.

In an earlier work [1] a dynamic mathematical model has been developed for the simulation of the dynamic behavior of the FCC pilot plant of CPERI. The riser simulator predicts the conversion, coke yield, and heat consumed by the catalytic reactions in the pilot riser on the basis of semi-empirical models developed in CPERI [2, 3]. The pilot regenerator model uses the two-phase theory of fluidization [4] with a dilute phase model in account for post-combustion reactions [5, 6]. The dynamic simulator has been validated against steady-state and dynamic experiments performed in the pilot plant [1].

For the case of the pilot plant, the variables that may be manipulated are the catalyst circulation rate and feed preheat temperature, whereas the effects of feedstock composition and catalyst quality are considered as disturbances. The controlled variables may be the riser temperature or/and the feed conversion. These two variables differ in nature regarding their controllability aspects. The riser temperature is an online measured variable, whereas the feed conversion is a variable that can only be measured after one hour of steady state plant operation. Additionally, constraints may be imposed on the CO, NOx, and SO2 emissions of the regenerator for their partial control. As the pilot plant regenerator operates under full combustion mode the goal of minimum CO emissions can be easily achieved, yet for the other goals, the potential of an optimal operating point of the pilot plant may be explored.

Typically, the MPC strategy involves specified prediction and control horizons. In such a scheme the optimal piecewise constant future control actions are calculated through dynamic optimization over specified time intervals. Multi-rate process measurement updates are obtained frequently, whereas the unobservable controlled variables (the feed conversion in this study) can be inferentially estimated using the dynamic model and the process measurements. The sensitivity of the performance of an optimal MPC structure in respect to the number of control intervals and the length of prediction and control horizons was examined in conjunction with the computational time necessary for the numerical solution. The MPC scheme allowed for an accurate targeting of the desired feed conversion with limited knowledge about the catalyst activity and selectivity, while showing robustness during simulation. In conclusion, the MPC strategy provided tight following of the prescribed operating conditions and objectives and promises the final reduction of redundant experiments with the same catalyst in catalyst evaluation tests.

Page 2: Model Predictive Control of a Fluid Catalytic Cracking

Introduction The control of industrial FCC units stands as one of the most challenging problems in

chemical process industry. This is due to the strong interactions and the high degree of uncertainty in the integrated riser-regenerator loop. The stochastic nature of the air distribution in the regenerator, the moderately defined flow regime of the gas-catalyst mixture in the riser and the catalyst circulation throughout the unit mainly contribute to the system uncertainty. Typical disturbances to the system are related to the different feed and catalyst qualities. Moreover, operational constraints that are set for safe (which typically means stable) operation, product specifications and environmental restrictions, further increase the challenges faced by the control problem.

The first necessity in the development of an efficient MPC scheme for the FCC is a detailed and accurate dynamic model. The dynamic simulation of the Fluid Catalytic Cracking (FCC) process is an intriguing research subject of high economic and environmental importance. Optimization of this complex process, demands the development of accurate models capable of describing the process in detail. On the other hand, the aforementioned uncertainty and the high degree of empiricism of the process operation, pose a limitation on model accuracy. In addition, the requirement for stable operation of commercial FCC units restricts the possibility of developing accurate models over an extensive operating range using experimental data, because the operating window is very narrow and the data is insufficient.

In industry, the maximum capacity (i.e. profitability) is the main target that restricts the application of dynamic transients within particular and narrow operating windows. Thus, FCC pilot plants are often used for obtaining data useful for the simulation of commercial units under different operating conditions, feed properties and catalyst activities and selectivities. The operation of a pilot unit provides the ability to examine the process under steady feed or catalyst properties, in order to isolate their respective effects on the cracking reactions and develop correlations for each subset of process variables. Another potential use of an FCC pilot plant is the examination of its dynamic behavior within a wide range of operating conditions, which enables the investigation of the process dynamics. Such a pilot plant operates in Chemical Process Engineering Research Institute (CPERI, Thessaloniki, Greece) and is mainly devoted in the performance of benchmarking experiments of new and promising catalysts under realistic conditions that closely simulate those of the refinery.

There are many similarities and several differences between the operation of the pilot plant and the one of a typical industrial unit. While the potential of yielding more market-oriented FCC products, increasing the capacity and stabilizing the operation are the typical concerns in industrial practice, the main concern in the pilot plant operation is a constant operation within a narrow predefined operational region. Therefore, control of this pilot scale process faces several challenges:

• riser temperature, affected notably by the catalyst circulation rate in the closed loop operation, should satisfy a specified set-point that guarantees constant selectivity in the product slate,

• conversion of the gas-oil feed should meet a determined value for easy comparison (testing) of variable catalyst activity and selectivity,

• flue gas from the regenerator is subject to environmental constraints regarding the CO, SO2, NOx emissions in commercial units and this pattern should be followed, or even examined, in the pilot plant operation, as well.

Page 3: Model Predictive Control of a Fluid Catalytic Cracking

Description and Modeling

Process description

The FCC pilot plant of CPERI (Figure 1) operates in a fully-circulating mode and consists of a riser reactor, a fluidized bed regenerator, a stripper and a liftline [7]. The riser reactor operates under pseudo-isothermal plug flow conditions and consists of a large-diameter bottom region (mixing zone) (26mm i.d., 0.3m height) and a smaller-diameter (7mm i.d., 1.465m height) top region connected by a frustum-shaped region 0.05m tall. At the reactor bottom, the gas-oil comes in contact with the hot catalyst (which flows in from the regenerator) and evaporates instantly, while the catalyst is preserved in a fluidized state by means of concurrent nitrogen flow. The cracked gas product from the riser top enters the stripper vessel for the separation (stripping) of vapors from the catalyst. Then the mixture flows through a heat exchanger for condensation of the heavier compounds. Thereupon, liquid and gas products are led to a cooled-top stabilizer column for separation. The mixture of gasoline, light cycle oil and heavy cycle oil is obtained through the bottom of the stabilizer. The liquid products yield is measured with the ASTM D-2887 simulated distillation method. The stripped catalyst flows through a liftline to the regenerator, where the carbon, deposited on the catalyst surface, is combusted. The regenerator consists of two main sections. A small-diameter bottom section (77.92mm i.d., 0.715m height) and a larger-diameter top section (254.6mm i.d., 0.64m height) connected by a frustum-shaped section 0.205m tall. A standpipe at the bottom of the regenerator leads the regenerated catalyst back to the riser bottom to continue the operation loop. The flue gas exits the system from the regenerator top.

Figure 1. Schematic diagram of the FCC pilot plant.

Two slide valves, one at the exit of the regenerator standpipe and one at the exit of the

stripper standpipe regulate the catalyst circulation throughout the unit. The regenerator standpipe slide valve controls the catalyst circulation to avoid aberrations from a predetermined set point for the riser temperature, while the stripper slide valve operates for constant stripper level (i.e. stripping volume). Two wet test meters and two gas chromatographs measure the volumetric flow rates and the molar composition of the flue and

Page 4: Model Predictive Control of a Fluid Catalytic Cracking

cracked gas, respectively. An on-line oxygen analyzer monitors the excess of oxygen and controls the combustion air flow rate to ensure thorough catalyst regeneration. The process control system of the unit is based on a special industrial computer control system which is coordinated by the iFIX S/W by Intellution GE. The control system collects the values of the inputs, drives the output signals and maintains a digital record of the signals, as well. The process pressure control valves and the power supply to electrical heaters are controlled by numerous algorithmic PID controllers.

Model description The simulator of the pilot plant includes three main sections: a pseudo-steady state model

of the riser reactor, a dynamic model of the regenerator and a set of dynamic and pseudo-steady state models of the stripper, the regenerator standpipe, the liftline and the slide valves. For the specific case of the CPERI pilot plant, the dynamic effects of the riser, the cyclones, the liftline and the regenerator standpipe were assumed negligible, as their operation has a significantly lower impact on the process dynamics, compared to the two large vessels of the plant, the stripper and the regenerator. The residence times of these two units are so much longer that the dynamic effect of the rest is suitably neglected. In both the pilot plant and in a typical commercial unit, the behavior of the regenerator dominates the dynamic and the steady state behavior of the integrated unit [8]. The riser residence times are much shorter compared to those of the regenerator, hence one can, at any instance, describe the riser reactor by a set of pseudo-steady state equations, which simplifies the dynamic analysis. Despite the pseudo-steady state assumption, the riser still affects the unit dynamics. The main impact of the riser operation on the dynamic (and steady state) behavior of the integrated system is through coke production and on heat consumption. That is because the amount of coke on spent catalyst and the catalyst rate entering the regenerator indirectly affect the dynamic behavior. For instance, a decrease on the regenerator temperature will result to a lower temperature stream entering the riser. The temperature drop is compensated by a rise of the catalyst circulation rate. With the catalyst circulation rising, the important side effect is the coke production increase of the riser, which in turn leads to a variation of the regenerator temperature (increase or decrease depending on the regenerator state), which again variably affects the riser and so on. Therefore, the accurate prediction of pseudo-steady state conversion, coke yield and heat of cracking and vaporization is significant, when describing the effect of riser in the integrated dynamic system.

The pseudo-steady state and dynamic sub-models that constitute the dynamic simulator of the pilot plant have been presented in the literature [2, 3, 5, 6] and are briefly presented in Appendices I and II. A coupled kinetic-hydro-dynamic model was developed for the simulation of the pilot riser reactor in pseudo-steady state conditions [2]. The catalyst hold-up and residence time in the reactor were calculated on the basis of empirical hydro-dynamic correlations and the gas-oil conversion and coke yield were predicted through a Blanding type [9] kinetic model. The prediction of gas-oil conversion and coke yield were the only two lumps of the riser sub-model essential for inclusion in the complete simulator, therefore using a more detailed lumped model was futile. The effect of feedstock properties on gas-oil conversion and coke yield was expressed through semi-empirical correlations developed on the basis of experiments performed with constant catalyst and a variety of feedstocks [3]. The effects of catalyst type and quality were expressed through a “catalyst index” [3]. This means that a different array of indices (one for each product) was assigned to each catalyst to express its activity and selectivity. The pseudo-steady state model of the FCC riser reactor [2, 3] was developed on the basis of the following assumptions:

• the aggregate effects of operating conditions, feed properties and catalyst type on the

Page 5: Model Predictive Control of a Fluid Catalytic Cracking

cracking reactions are simulated by the product of their discrete functions • the riser reactor is assumed to be running in concurrent plug flow of gas and solids

under pseudo-isothermal conditions • second-order rate apparent kinetics are applied for gas-oil conversion (yx) • catalytic coke (

xcy ) deposition parallels catalyst deactivation [10]. The model of the regenerator was based on the two-phase theory of fluidization [11], in

which the gas-solids flow is assumed to follow the bubbling bed regime, consisting of two zones: a dense zone at the regenerator bottom comprised by a bubble and an emulsion phase, and a dilute zone at the regenerator top, called the freeboard. The assumptions made for the simulation of each phase [5, 6] are:

• the bubble phase is free of catalyst particles • plug flow regime is assumed for the bubble phase • the emulsion phase gas and catalyst particles are assumed fully mixed • the freeboard is modeled as an ideal plug flow reactor • the catalyst particles are hydro-dynamically represented by their average size,

density and porosity, while the particle size distribution is used for the emulsion to freeboard entrainment rate calculation

• diffusion in the catalyst particles is neglected • due to the high temperatures in the FCC regenerator, the ideal gas law is valid • the regenerator reactor is adiabatic. The model equations were grouped into two main modules that serve for the two main

sections of the unit, the riser and the regenerator. A third module was used for the simulation of the stripper and the slide valves, the liftline and the standpipe. The assumptions made for the simulation of the stripper are:

• the stripper is a perfectly mixed reactor in minimum fluidization conditions • the stripping efficiency of the pilot stripper is 100%. The dynamic material and energy balance equations form a system of Differential-

Algebraic Equations (DAE) that was initially developed and adapted in FORTRAN programming language. However, the dynamic optimization algorithms in a typical MPC scheme require a frame developed in open form (equation based form) [12], which offers better flexibility. Thus, the following step was the integration of the model in the equation based environment of gPROMS [13]. Finally, the model and the MPC algorithm were merged in a compound module, forming a fully operational routine which was used for the purposes of this project.

Model-based control

Control Objectives

The main task of the pilot plant is catalyst benchmarking. The interest is in maintaining the operation within a predefined span in order to achieve conversion values that are considered as standard by the industry. This practice is especially adopted for gathering comparable results in terms of catalyst selectivity or even feed quality effect so any experiment not fulfilling that requirement is useless. In this way, the overall control objective translates to the elimination of repetitive and useless experiments.

The initial approach presented in this project will focus on improving the control performance of the unit through the manipulation of the riser. That is dictated by the operational conditions of the CPERI pilot regenerator, which allow a small margin for

Page 6: Model Predictive Control of a Fluid Catalytic Cracking

optimization, since the primary target for minimal polluting emissions is in any case achieved through full combustion of the coke under excess air conditions. So, as a first step, the optimal control case will be explored without imposing additional constraints (minimal CO emissions or specified regenerator temperature) considering the regenerator operation. Still, the regenerator operation significantly interacts with the riser. As mentioned in the previous section, the regenerator defines the dynamics of the unit. MPC optimal control of the riser cannot be achieved via a stand-alone riser model. It requires an integrated accurate model of the riser-regenerator system, mainly due to the fact that both vessels interact with each other through stream recycling. As seen in Figure 1, any variation in the coke/catalyst output stream of the riser is eventually led to the regenerator. The regenerator operation in turn, is notably affected by the transition of the input, resulting in dynamically changing operating conditions and output composition, which are fed back to the riser as a recycle stream affecting it again. It is obvious that the operation of the regenerator must not be neglected and that modeling and monitoring the regenerator is essential for controlling the riser. The effect of the accuracy of the model on the controlling efficiency is discussed by the Model Predictive Control theory [12]. In general, the model of the process should be accurate enough to maintain good prediction properties over the range of operating conditions of interest. The model used for this project has been developed and presented in detail in other relevant publications [1], in which its ability to simulate accurately the dynamics of the pilot plant has been demonstrated.

A robust control system presupposes a suitable choice of the controlled (output) and manipulated (input) variables. The manipulated variables should be the ones that highly affect the unit outputs. Moreover, they must allow for operational flexibility and posses the ability to alleviate the effects of disturbances successfully. The controlled variables should be chosen according to the control objectives (yield maximization, temperature stability etc.). At this point, it is imperative to present a brief degrees of freedom analysis of the system. The riser sub-section is described by 5 basic equations: ( ) ( )( )RS, , , , , ,x x F C N RXy f W W W P T p feed p cat= & & & (1)

( ) ( )( )RS, , , , , ,c c F C N RXy f W W W P T p feed p cat= & & & (2)

( )( ) ( ) ( )( )

( )( ):RG

:RS

, , , ,

, , 0

tF F PR RX C C D RX

N N D RX vap crack

W f T T p feed W f T T p cat

W f T T p inert H H

× + × +

× + ∆ + ∆ =

& &

& (3)

( )( ):RS, , , , ,vap F v F C PR RX DH W f W W T T T p feed∆ = ×& & & (4)

( )( ), ,crack F k x RXH W f y T p feed∆ = ×& (5) Practically, the system of eqs.(1) - (5) briefly describes the mass and energy balances of

the riser section. More details on those are provided in the Appendices (eq.(12) - (15)). The operational variables (unknowns) occurring above, are 14: the conversion (yx) and the coke yield (yc), the feed rate ( FW& ) the inert rate ( NW& ), the catalyst circulation rate ( CW& ), the pressure (

RSP ) and the temperature ( RXT ) of the reactor, the feed preheat temperature ( PRT ),

the temperature at the regenerator dense section ( ( ):RGt

DT ), the feed vaporization enthalpy ( vapH∆ ), the energy consumed by the cracking reactions ( crackH∆ ) and the properties or quality indices of the feed (p(feed)), the inert (p(inert)) and the catalyst (p(cat)). The respective schematic system of equations for the regenerator (neglecting the dynamic clauses of the stripper) are the following: ( ) ( ) ( ) ( ) ( ) ( )( )0 0 1,

:RG :RG :RG :RG :RG :RG RG, , , , , , , ,D D Ft l l l t t tc c c F g g g c C Dc f y W W c c c W T P= = == & & & (6)

( ) ( ) ( ) ( ) ( ) ( ) ( )( )1, 0 0 1, :RG :RG :RG :RG :RG :RG RG, , , , , , , , ,F D D Fl t l l l t t t

g g c F g g g c C Dc f y W W c c c W T P p air= = = == & & & (7)

Page 7: Model Predictive Control of a Fluid Catalytic Cracking

( ) ( )( ) ( ) ( ) ( ) ( ) ( )( )0 0 1:RG :RG :RG :RG :RG, , , , , 0D D Ft l l t l

C C RX D g g g D F combW f T T p cat W f T T T p air H= = =× + × + ∆ =& & (8)

( ) ( ) ( ) ( ) ( )( )0 0 1:RG :RG :RG :RG RG, , , , , , , ,D D Fl l t l

comb r c F g C g D FH f y W W W T T T P p air= = =∆ = & & (9)

The system of eqs.(6) - (9) describe very briefly the extended system of equations of the regenerator. The variables appearing here are 8: the air (for combustion) flow rate at the bottom of the regenerator ( ( )0

:RGDl

gW =& ), the composition of the gas exiting from the regenerator

top ( ( )1, :RGFl t

gc = ), the composition of coke on the regenerated catalyst ( ( ):RGt

cc ), the average

regenerator pressure ( RGP ), the inlet temperature of the combustion air ( ( )0:RG

DlgT = ), the heat of

the exothermic reactions ( combH∆ ) the combustion air composition ( ( )0:RGDl

gc = ) and its properties (p(air)). The system in its general form totals 22 variables and 9 equations, therefore 13 degrees of freedom. It is noted that the analysis presented, despite the generalities, requires full comprehension of the FCC unit operation and it can be easily adopted by any pilot or industrial unit.

The 13 independent variables of the FCC operation are described below: the feed rate ( FW& ), the inert rate ( NW& ), the riser pressure ( RSP ), the catalyst circulation rate ( CW& ), the feed preheat temperature ( PRT ) and the qualities of the feed (p(feed)), the inert (p(inert)) and the catalyst (p(cat)) for the riser section. For the regenerator they are the air rate at the bottom ( ( )0

:RGDl

gW =& ), the pressure ( RGP ) and the inlet air temperature ( ( )0:RG

DlgT = ), composition ( ( )0

:RGDl

gc = ) and properties (p(air)). In a typical industrial unit: 1) the feed rate is set to meet the maximum capacity of the unit, 2) the inert rate and 3) the inert quality follow a predefined pattern in order to retain constant partial pressure of the hydrocarbons in the riser, 4) the temperature, 5) the composition and 6) the properties of the combustion air are constant. These bounds are also followed in pilot plant operation, although not always for the same reasons. Moreover, specifically in the pilot plant: 7) the combustion air rate of the pilot regenerator is controlled separately in order to satisfy the low emissions criterion. Also, 8) the feed quality and 9) the catalyst quality are considered as unknown disturbances. The reason for the latter is their stochastic nature in industry, meaning that the feed quality description is usually unavailable, because it is a mixture of various refinery streams and the catalyst quality is changing perpetually due to the continuous addition of a small amount of fresh catalyst. Furthermore, feed or/and catalyst qualities are the usual unknowns during pilot plant benchmarking experiments. Provided that 10) the riser pressure and 11) the regenerator pressure are controlled by separate subsystems, the only independent variables suitable for manipulation in the pilot plant, for the purpose of benchmarking experiments, are the catalyst circulation rate ( CW& ) and the feed preheat temperature ( PRT ). The catalyst circulation rate is always chosen for manipulation, but regarding the feed preheat temperature, two cases may be explored. It may be treated equally to the catalyst rate as a manipulated variable or it may be set constant to match the optimal value of the one in an industrial unit.

The main objective in both the pilot plant and the industry FCC process is the optimization of the riser conversion on feed basis, while maintaining the riser temperature around a set point, which guarantees the constant effect of operating conditions on product selectivity. In industry the conversion control targets maximum profitability. In the pilot plant, where the intention is the efficient catalyst benchmarking, the interest lays in maintaining the operation within a predefined span. The riser conversion, the riser temperature, the feed preheat temperature and the catalyst circulation rate are interrelated variables and comprise a system of equations (eqs.(1) - (5)), which under stable operation is uniquely defined (within the narrow bounds of the pilot plant operation). The manipulated

Page 8: Model Predictive Control of a Fluid Catalytic Cracking

catalyst circulation rate obviously affects the conversion, but it also affects the heat build-up, consumption and loss of the system so it has an impact on the riser temperature. Riser temperature and the riser conversion are correlated. This implies that for a given value of the riser temperature, conversion is defined uniquely and vice-versa (given that every other input variable of the riser is constant). The last fact provides two alternatives for the control problem: If riser conversion measurements were available then it might be directly controlled by manipulating the catalyst rate and feed preheat. In the most usual case that the riser temperature measurement is more accessible and no conversion measurement is available online, then conversion control can be performed using an inferred value calculated by eq.(1).

On the basis of the above analysis, an MPC strategy can be implemented for the control of conversion on feed basis and riser temperature through the proper manipulation of the catalyst circulation rate and the feed preheat temperature. The strategy provides full automation and control of the riser section performance. Furthermore, the direct targeting of the desired conversion reduces significantly the number of required experiments with the same catalyst in catalyst evaluation tests.

Model-based predictive control

Model predictive control is based on the fact that past and present control actions affect the future response of the process [12]. Having selected a time horizon extending into the future, the prediction of the process model is calculated, on the basis of the past control actions. The response of the model can then be compared to a desired trajectory if no further control actions are to be taken. The difference between the desired control trajectory and the predictions can therefore be minimized, through the calculation of a specified number of future control actions (Figure 2). The control horizon (i.e. the period for which future control actions are calculated) may be selected smaller or equal to the prediction time horizon, during which the comparison of the desired to the predicted trajectories is performed. At each time interval the first optimal control action in the calculated sequence is implemented and a new measurement of the actual response of the process is obtained. The model-based predictive control principles are presented in Figure 2.

Desired trajectory

Predicted trajectory

ek+2

Rolling time horizon

uk+1 uk+2 uk+3uk uk-1 uk-2

Past control actions Calculated control actions tk+1 tk+2 tk+3 tk tk-1 tk-2 tk+4

Figure 2. Schematic representation of the principles in model-predictive control.

The difference between the actual response of the process and the model prediction is

recorded and considered as the error of the process model, as shown in the block diagram of

Page 9: Model Predictive Control of a Fluid Catalytic Cracking

the MPC system (Figure 3). The calculated error defines a bias term that is used to correct the future predictions of the model. The bias model term encompasses contributions from model mismatch, unmeasured disturbances, and measurement error. It is assumed that the error calculated in the behavior of the system would be persistent for the entire prediction horizon. Therefore, error feedback is maintained in the control system that allows for integral action and the elimination of steady-state offset. The block diagram describing the system is presented in Figure 3. Optionally, a parameters and state estimator can be added to enhance the model accuracy and the overall MPC robustness.

The mathematical representation of the model-based predictive control algorithm is given by the following expression:

( )( )

( )

( ) ( )

1 11

2 2 2

1 1 11 1 1

1 1

1

1

ˆmin

subject to:

ˆ

/ , /

C CP

y u uk j k jk jk j

N NNsp ss

MPC k j k j k j k j k jj j j

meas predk j k j

predk j k j k j

l uk j

C C k C P P k PN T T t N T T t

∆+ − + −++ −

+ + + − + − + −= = =

+ − + −

+ + + −

+ −

= − + + −

= −

= +

≤ ≤

= − ∆ = − ∆

∑ ∑ ∑w w wuy y ∆u u u

x = f x,u

y = g x,u

e y y

y y e

u u u

J

&

(10)

where x, u, y denote the vectors of the state, manipulated (i.e. control actions) and output variables of the system, respectively. Symbols f and g denote the sets of differential and algebraic model equations. Vector y denotes predictions for the system output variables that include the contribution of the bias term on the model predictions. Vector ysp denotes the desired response (set point) of the system output variables. Vector ek denotes the difference between the measured output variables ymeas and the predicted values ypred at time instant k. The current formulation assumes that the error on the predictions will persist and remain constant for the entire length of the prediction time horizon. TP and TC denote the prediction and control horizons, reached through NP and NC time intervals, respectively.

MPC PROCESS

MODEL

ESTIMATOR

manipulatedvariables

(u)

measuredoutputs(yi

meas)

predicted outputs(yi

pred)

states(x)

modelparameters

(θ)

desired trajectory

(yisp)

disturbance(dk)

error(ek)

outputvariables

(yi)

Figure 3. Control block diagram of the process.

The tuning parameters of the controller are the weights wy, wu and w∆u, and the length of

the prediction and control horizons. A long prediction horizon allows the control scheme to

Page 10: Model Predictive Control of a Fluid Catalytic Cracking

compensate for slower dynamics that affect the response of the system further into time. However, large prediction horizons make the control scheme more susceptible to unmeasured disturbances. On the other hand, a short control horizon may lead to aggressive control actions, as the controller attempts to correct the trajectory with a few moves through short time period.

Solution method

The formulation of the control problem (eq.(10)) results in a dynamic program. The objective function contains the integral of the squared error of the controlled variables from the desired trajectory, a move suppression factor on the manipulated variables that penalizes high values in the rate of change for the control actions and a steady state optimality factor for restricting the range of the possible solution within the operational limits. The behavior of the manipulated variables is considered as a sequence of piecewise constant values that minimize the value of the objective function. The prediction and control horizons are divided in equally spaced time intervals in which the level of the manipulated variables remains constant. Upper and lower bounds apply for the manipulated variables along the control horizon, as required by the physical limitation of the system (e.g., CW& cannot exceed its value for the respective maximum available valve opening or the minimum flow necessary for safe operation). The modeling equations are satisfied along the trajectory of the system. The solution method involves successive iterations between an optimizer that evaluates the optimal values for the manipulated variables and an integrator that calculates the dynamic behavior of the system and the sensitivity of the control actions to the control objectives. Variable bounds and path constraints are considered for violation along the optimal control path [13], [14].

Results

Model-based predictive control of FCC reactor (simulated results)

The model predictive controller was initially tested on a simulated case study. The MPC framework includes two instances of the model that were concurrently executed. The first instance, which represented the “Virtual Process” or “Virtual Plant” (VP), was depicted by a flawless version of the model. The second introduced a case study including significant amount of mismatch in the reaction kinetics in order to simulate a fictitious simulated process and was used to represent the “Simulator”. Hence, in the following the expression “Virtual Process” denotes the process, for which the flawless version of the model was used, and the expression “Simulator” denotes the model with the different kinetic constants. This case study actually included what is expected to be the control problem in the real pilot process level. More specifically, different indices that describe the effect of catalyst activity and selectivity on feed conversion and coke yield were used in the “Virtual Process” model and the “Simulator” model. The indices used were those of real catalysts, different for each case (“Virtual Process” and “Simulator”) taken from the experimental database of the pilot plant. With this structure the equivalence of a typical catalyst benchmarking experiment was fully reconstructed. The intervals for the control actions (i.e. manipulation of variables) were chosen equal to 2 minutes. The “Simulator” was updated using infrequent rate process measurement of the reactor temperature and the inferred conversion of the virtual process. The optimal piecewise constant future control actions were calculated through dynamic programming over the desired prediction and control horizons.

Page 11: Model Predictive Control of a Fluid Catalytic Cracking

Validation of the model-based predictive control scheme

The different catalyst activity and selectivity between the “Virtual Plant” and the “Simulator” act as a constant disturbance in the process cycle. The goal for the MPC was to move the plant through a sequence of corrective control actions to the desired level of feed conversion. The controlled variables, namely the feed conversion and riser temperature should remain as close as possible to the set-point value. The following performance index was used (eq.(11)):

( ) ( )( ) ( )12 2

2/ 1

1 1 1

ˆ ˆˆ1 1

y u uk

k

tn n ni i ky u u

k i i i k k isp ssi i ii it

y t u tJ w w u t w

y u

+∆

−= = =

⎛ ⎞ ⎛ ⎞= − + ∆ + −⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠∑ ∑ ∑∫ (11)

where ˆiy denotes the predictions of the respective variables i that incorporate the model prognosis and the error correction (i.e. difference between the measurement and the prediction at the previous time period) and sp

iy the set points of the controlled variables. The second term of eq.(11) denotes the move suppression factor, which penalizes abrupt changes in the manipulated variables. The last term denotes the steady state optimality factor and it aims to drive the MPC actions towards a potentially desired solution, dictated by the plant optimization decision level. Weights w express the relative significance of each term in the performance index.

The prediction (NP) and control (NC) horizons were selected equal to 20 and 10 minutes, respectively. The length of the prediction horizon is close to the time necessary for the pilot plant to reach the new steady state after imposing a typical change. The length of the control horizon was mainly driven by the computational time for solution that should be lower than the unit sampling interval. The control profile was considered as piecewise constant with the manipulated variables changing every 2 minutes. The length between two consecutive control (∆tC) actions was chosen based on the frequency of the available measurements. A new optimal sequence of 5 control actions was calculated every 2 minutes. This means that every 2 minutes a new control action was implemented and a new measurement was recorded. The time between two successive measurements was determined considering also the limitation imposed from the computation time required for the solution of the dynamic optimization and simulation of the process model.

The control problem, as posed above, was tested on a simulation environment, in which different catalyst indices (kx, kc) were used for the virtual process and the simulator model. The indices of a catalyst with higher activity and selectivity and much higher coke selectivity were used in the virtual process. These significant differences caused an increase in the feed conversion of the virtual plant and lower riser temperature, compared to those predicted by the simulator model. The task for the MPC algorithm was to lead the virtual process to the desired conversion ( 65 %wtsp

xy = ) and riser temperature ( 526.7 CspRXT = ° ) under the

influence of the disturbances generated by the catalyst. A bias term (ek of eq.(10)) was used for the update of the process model. Although it is known that a constant bias term (referred also as constant additive disturbance model [14]) is not the best practice to compensate model error, it will be shown that the proposed structure handled the model mismatch robustly. A module that will perform state and parameters estimation will be developed in a following stage of this project.

As shown in Figure 4, the “Virtual Process” initiated at 1% higher feed conversion than the desired one and riser temperature 1% above its set point. The first action of the MPC was to lower the catalyst circulation rate and increase the feed preheat temperature and, as dictated by the solution of the dynamic problem. The lower catalyst circulation rate led to lower coke yield (on feed basis), but higher overall ratio of coke rate over catalyst rate entering the regenerator. The latter resulted in increasing the regenerator temperature (Figure 4(f)) and

Page 12: Model Predictive Control of a Fluid Catalytic Cracking

eventually the riser temperature. This is evident by the curve of the riser temperature in the first set of control actions (Figure 4(d)). The MPC using the information of the prediction horizon led the “Virtual Process” to a state of lower riser temperature and then waited for the dynamics of the process, while making small control actions to compensate for the “Virtual Process” – “Simulator” mismatch. The simulated control loop was continued for a period of 30 min. In the final steady state both the feed conversion and the riser temperature criteria were fully satisfied. In Figures 4(g) and (h) it is evident that the regenerator flue gas is not significantly influenced by the sequence of control actions, because of the large excess air supply. The operation in full combustion mode is the reason for the practically zilch CO and the very low SO2 emissions. Overall, the MPC structure presented promises the establishment of the desired steady state within 40 min, which is very important for the pilot plant operation.

Figure 4. Results of the application of the MPC structure on a simulation environment

(continuous lines denote the “Virtual Process” and dotted lines the “Simulator”).

Page 13: Model Predictive Control of a Fluid Catalytic Cracking

Conclusions An advanced model predictive control strategy that calculates the optimal sequence of

manipulated variables over a specified control horizon has been implemented in a pilot- sized FCC unit used for catalyst experimentation. The implementation of the MPC scheme showed extreme robustness to changes in the catalyst activity and selectivity. The MPC scheme allowed for an accurate targeting of the desired feed conversion with little knowledge about the catalyst properties and selectivity. In conclusion, the MPC strategy allowed for tight following of the prescribed operating conditions and the elimination of additional and/or repeat experiments with the same catalyst in catalyst evaluation tests, thus improving the overall productivity of the catalyst evaluation studies task that the pilot plant is mainly used for.

Acknowledgment Financial support by the European Social Fund & National Resources - EPEAEK II –

ARCHIMIDIS is gratefully acknowledged.

References

1. Bollas, G.M., et al., Integrated FCC Riser - Regenerator Dynamics studied in a Fluid Catalytic Cracking Pilot Plant. Chemical Engineering Science, 2006. (submitted for publication).

2. Bollas, G.M., et al., Modeling small-diameter FCC riser reactors. A hydrodynamic and kinetic approach. Industrial & Engineering Chemistry Research, 2002. 41(22): p. 5410-5419.

3. Bollas, G.M., et al., Bulk Molecular Characterization Approach for the simulation of FCC feedstocks. Industrial & Engineering Chemistry Research, 2004. 43(13): p. 3270-3281.

4. Kunii, D. and O. Levenspiel, Fluidization Engineering. 1977, Florida: Robert E. Krieger Publishing Company Inc.

5. Faltsi-Saravelou, O. and I.A. Vasalos, Fbsim - A model for fluidized-bed simulation. 1. Dynamic modeling of an adiabatic reacting system of small gas-fluidized particles. Computers & Chemical Engineering, 1991. 15(9): p. 639-646.

6. Faltsi-Saravelou, O., I.A. Vasalos, and G. Dimogiorgas, Fbsim - A model for fluidized-bed simulation. 2. Simulation of an industrial fluidized catalytic cracking regenerator. Computers & Chemical Engineering, 1991. 15(9): p. 647-656.

7. Vasalos, I.A., et al. Design Contsruction and Experimental Results of a CFB FCC Pilot Plant. in Circulating Fluidized Bed Technology V 1996. Beijing, China: Science Press.

8. Arbel, A., et al., Dynamics and control of Fluidized Catalytic Crackers. 1. Modeling of the current generation of FCCs. Industrial & Engineering Chemistry Research, 1995. 34(4): p. 1228-1243.

9. Blanding, F.H., Reaction rates in the catalytic cracking of petroleum. Industrial & Engineering Chemistry, 1953. 45: p. 1186.

10. Voorhies, A., Carbon formation on catalytic cracking. Industrial & Engineering Chemistry, 1945. 37: p. 4.

11. Davidson, J.F., R. Clift, and D. Harrison, Fluidization. 1985, London: Academic Press Inc.

Page 14: Model Predictive Control of a Fluid Catalytic Cracking

12. Marlin, T.E., Process Control: Designing Processes and Control Systems for Dynamic Performance. 1995, New York: McGraw-Hill Inc.

13. Process Systems Enterprise, P.S.E., gPROMS Advanced User’s Guide. 2004: London, UK.

14. Nagy, Z.K., et al. A real-time approach for moving horizon estimation based nonlinear model predictive control of a fluid catalytic cracking unit. in 7th World Congress of Chemical Engineering, GLASGOW2005, incorporating the 5th European Congress of Chemical Engineering. 2005.

15. Kesler, M.G. and B.I. Lee, Improve prediction of enthalpy of fractions. Hydrocarbon Processing, 1976. 55(3): p. 153-158.

16. Richardson, J.F. and W.N. Zaki, Sedimentation and Fluidization. I. Transactions of the Institution of Chemical Engineers, 1954. 32: p. 35.

17. Pugsley, T.S. and F. Berruti, A predictive hydrodynamic model for circulating fluidized bed risers. Powder Technology, 1996. 89(1): p. 57-69.

18. Abrahamsen, A.R. and D. Geldart, Behaviour of gas-fluidized beds of fine powders, part I. Homogeneous expansion. Powder Technology, 1980. 26(1): p. 35-46.

19. Geldart, D., Elutriation, in Fluidization, J.F. Davidson, R. Clift, and D. Harrison, Editors. 1985, Academic Press Inc.: London. p. 383.

20. Patience, G.S., et al., Scaling considerations for circulating fluidized bed risers. Powder Technology, 1992. 72(1): p. 31-37.

21. Judd, M.R. and P.D. Dixon, Flow of fine, dense solids down a vertical standpipe. AIChE Symp Ser, 1978. 74(176): p. 38-44.

APPENDIX I: Mathematical Model Formulation

Simulation of Riser

On the basis of the assumptions mentioned above and after integration and rearrangement of the corresponding spatial equations eqs.(12) and (13) were formulated:

( ) ( ) :RS0 exp100

xnx x xC

g RX x

k E yC catalyst type F feed quality tWHSV R T y

⎛ ⎞−= −⎜ ⎟⎜ ⎟ −⎝ ⎠

(12)

( ) ( ) :RS0 exp c

x

nc cc c C c

g RX

k EC catalyst type F feed quality t yWHSV R T

⎛ ⎞−= −⎜ ⎟⎜ ⎟

⎝ ⎠ (13)

The main contributors to the overall enthalpy balance in an FCC plant are the enthalpy of cracking ∆Hcrack; the enthalpy of vaporization of the gas-oil feedstock; and the enthalpy content of various process streams (gas-oil, catalyst, cracked products, inerts), as shown in eq.(14): -0 crack vap gas oil cat LossH H H H H= ∆ + ∆ + ∆ + ∆ + ∆ (14)

The enthalpy content of gas-oil vapors was estimated by integration of the empirical correlation of Kesler and Lee [15]. The heat of cracking was estimated as a function of conversion, riser temperature and gas-oil molecular weight, as shown in eq.(15):

( ) ( )2 21 2 3 1 2 30 ln

100 RX RX F RX RX F crackx a T a T a MW b T b T b MW H

x⎛ ⎞= + + + + + − ∆⎜ ⎟−⎝ ⎠

(15)

Simulation of Regenerator

Page 15: Model Predictive Control of a Fluid Catalytic Cracking

The dense bed volume was calculated on the basis of the overall regenerator dynamics:

( ) ( ) ( ) ( )

( )1 0 1 0

:RG :RG :RG :CY :CY

1

D D F Fl l l lD C C C C

p e e

dV W W W Wdt fρ ε

= = = =− + −=

& & & & (16)

The material balance for gas components in the bubble phase is:

:RG

1 homoib

Mi b ij RjbjD D

dF K f a KV dl

= − + ∑ (17)

The energy balance in the bubble phase is given by eq.(18):

( ):RG

1 homob

H b Rj RjbjD D

dQ K f H KV dl

= − + −∆∑ (18)

In the emulsion phase the material balance equations were formulated separately for gas and solids components, as shown in eqs.(19) and (20) respectively:

( ) ( )

( )0 10

:RG 0

1D Dl l homo hete

geie ie iee e Mi D e e ij Rje e e ij Rje

j jge D

Wdc c cf K dl f a K f a Kdt V

ε ε ερ

= = −= + + + −∑ ∑∫

& (19)

( )( ) ( ) ( ) ( )

( )11 1 1

:RG :CY

:RG :RG

1 1FD D F ll l l hete

if ieie C ie ie Ce e e e ij Rje

jp D p D

c cdc W c c Wf f a Kdt V V

ε ερ ρ

== = = −−− = + + − ∑

& & (20)

The energy balance equation in the emulsion phase is given by eq.(21):

( ) ( ) ( ) ( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

1 0 1 0 0:RG:RG :RG :CY :CY

11

:RG :RG :RG0

1

1

D D F F D

D

gassolidsl l l l lD e

e e ie ie e e ie ie C C C C gei i

homo hetel

ge loss D H D e e D Rj Rje e e D Rj Rjej j

d V Tf c cp f c cp Q Q Q Q Q

dt

Q Q V K dl f V H K f V H K

ε ε

ε ε

= = = = =

=

⎛ ⎞− + = − + − +⎜ ⎟

⎝ ⎠

− − + + −∆ + − −∆

∑ ∑

∑ ∑∫(21)

The superficial bubble gas velocity for the dimensionless fraction of dense bed height lD, is evaluated by differentiating the ideal gas law in terms of the bubble enthalpy rate term:

:RG :RG

gb g b

D D D gb D

du R dQdl A P cp dl

= (22)

The bubble-emulsion mass interchange MiK and the heat interchange HK and the emulsion fraction ef are evaluated by eqs. (23) - (25), respectively:

:RG

0 ibti ie Mi

gb D

FK c Ku A

⎛ ⎞= − −⎜ ⎟⎜ ⎟

⎝ ⎠ (23)

( )0 t b e HH T T K= − − (24)

( )1

0

1e b Df f dl= −∫ (25)

The combined bubble to emulsion gas interchange coefficients are evaluated by eq.(26):

1 10 b

bci cei ti

fk k K

⎛ ⎞= + −⎜ ⎟⎝ ⎠

(26)

For the evaluation of the bubble-cloud (kbci) and cloud-emulsion (kcei) gas interchange coefficients the expressions proposed by Kunii and Levenspiel [4] were adopted. The same method was used for the estimation of the heat interchange coefficient (Ht) [4].

The freeboard is simulated as an ideal two-phase PFR. The material balances of the gas and solid components in the freeboard are shown in eqs. (27) and (28), respectively:

( ):RG

1 1homo hete

iff ij Rjf f ij Rjf

j jF F

dFK K

V dlε α ε α= + −∑ ∑ (27)

Page 16: Model Predictive Control of a Fluid Catalytic Cracking

( ):RG

1 1hete

iff ij Rjf

jF F

dFK

V dlε α= − ∑ (28)

The energy balance for the freeboard is:

( ) ( ) ( ):RG

1 1homo hete

ff Rj Rjf f Rj Rjf

j jF F

dQH K H K

V dlε ε= −∆ + − −∆∑ ∑ (29)

The ideal gas law is differentiated in terms of the gas enthalpy rate to evaluate the gas superficial velocity:

:RG :RG

gf g gf

F F F gf F

du R dQdl A P cp dl

= (30)

The derivative of the enthalpy of the gas phase is obtained by eq.(31), assuming that the heat capacity of the components is constant at each integration step:

gf gf f

F f F

dQ Q dQdl Q dl

= (31)

Simulation of Stripper and Slide Valves

The stripping volume and the material balance for the solids components were expressed through eqs.(32) and (33):

( ) ( )

( )1 0

:ST :ST :ST

1

D Dl lD C C

p mf

dV W Wdt ρ ε

= =−=

& & (32)

( ) ( )( )( )

1 1:ST :ST :ST:ST

:ST1

D Dl lC i ii

p mf D

W c cdcdt Vρ ε

= = −=

& (33)

The temperature of the catalyst stream at the exit of the regenerator standpipe (riser entrance) and at the exit of the liftline (regenerator entrance) was calculated by modeling the heat loss throughout their height with eq.(34):

( ) w w ww

C C

U D LdT T Tdl W cp

π= −

& (34)

The catalyst with concentration ( )tiec enters the riser with rate that is determined by the

slide valve at the end of the regenerator standpipe (eq.(35)):

( ) ( ) ( )( )( )1

1

1

0.52 2 0.5SP :SV 1 1

:RS SV RG RG SP RS2 2SP :SV

2 1 F Ft l lC p b RS

t

A AW k P P P P P

A Aρ ε = =⎛ ⎞

= − + ∆ + ∆ − −∆⎜ ⎟⎜ ⎟−⎝ ⎠& (35)

The same formulation is used to calculate the catalyst mass flow rate entering the regenerator:

( ) ( ) ( ) ( )( )( )2

2

2

0.52 2 0.5ST :SV1 1 1

:RG SV ST ST RG LL2 2ST :SV

2 1D F Ftl l lC p mf

t

A AW k P P P P

A Aρ ε= = =⎛ ⎞

= − + ∆ − −∆⎜ ⎟⎜ ⎟−⎝ ⎠& (36)

APPENDIX II: Hydrodynamic Correlations and Pressure Balance

Simulation of Riser

The weight hourly space velocity (WHSV) and the solids residence time (ts) were

Page 17: Model Predictive Control of a Fluid Catalytic Cracking

calculated by eqs.(37) and (38), following the pilot riser geometry, that is divided in three regions:

( ) ( ) ( )( )

:RS

:RS :RS :RS :RS :RS :RS

036001 1 1

F

p D D C C F F

W WHSVV V Vρ ε ε ε

= −− + − + −

& (37)

:RS:RS

:RS

36000

FC

C

W tWHSV W

= −&

& (38)

(a) The mixing region at the riser bottom. The void fraction ( :RSDε ) and subsequently the catalyst inventory of this region, were related to the superficial gas velocity by means of the empirical correlation of Richardson and Zaki [16] (eq.(39)), which substantiates for a dense regime in the bottom region of the pilot unit.

1

:RS:RS

:RS

0/ z

g:DD

t

uu

ε⎛ ⎞

= −⎜ ⎟⎝ ⎠

(39)

(b) The conical shaped intermediate region. Because of the very small volume of the intermediate region (15% of total riser volume), a simple approximation of averaged (between top and bottom regions) hydrodynamic attributes was used [17].

(c) The fast fluidization region at the riser top, which was simulated under the following assumptions: (i) the flow is fully developed, thus its hydrodynamic features remain constant with height; (ii) the total volumetric yield of the reaction is flowing through the whole height of this region; (iii) the particle acceleration is considered to be negligible. Hence, eq.(40) holds:

: :RS :RS:RS

:RS :RS : :RS :RS

0 g F p FF

F C g D p F

u Ay W u A

ερ

ρ= −

+& (40)

In eq.(40) :RSFy is the average gas-solids slip factor for the top section of the riser, which was proven to play an important role in small diameter riser reactors [2]. The correlation of Pugsley and Berruti [17] was applied for the estimation of the gas-solids slip factor as shown in eq.(41), where Frg and Frt are the Froude numbers for the superficial gas velocity and solids terminal velocity, respectively:

0.41: :RS :RS2

: :RS

5.60 1 0.47 t F Fg F

Fr yFr

= + + − (41)

A detailed pressure gradient analysis is required for small diameter risers [2]. For this analysis, all pressure gradients should be taken into account, and eq.(42) is valid where ∆Pfg is the gas-wall frictional pressure drop, ∆Pfs is the solids-wall frictional pressure drop, ∆Pacc is the pressure drop due to solids acceleration, and the other terms represent the pressure drop due to solids and gas static head throughout the total riser height: ( )RS :RS :RS :RS RS :RS RS RS RS1fg fs acc g pP P P P gL gLε ρ ε ρ∆ = ∆ + ∆ + ∆ + + − (42)

Simulation of Regenerator

For group A particles the emulsion gas superficial velocity is the gas velocity for zero net flow of solids, which equals the minimum bubbling velocity, plus (concurrent gas/solids flows) or minus (countercurrent gas/solids flow) the superficial solids velocity in the emulsion phase: 0 mb se geu u u= ± − (43)

For the evaluation of the minimum fluidization velocity the equation of Wen and Yu [4] is applied. For group A particles the minimum bubbling velocity, umb, is evaluated by the

Page 18: Model Predictive Control of a Fluid Catalytic Cracking

correlation of Abrahamsen and Geldart [18], which considers the effect of catalyst fines, f , on umb:

( )

( )0.126 0.523

0.9340.8 0.934

2300 exp 0.7160 ge ge mb

mfp p ge

uud g

ρ µ

ρ ρ= −

f (44)

The superficial gas velocity in the dense zone is then obtained by eq.(45): :RG0 gb ge gu u u= + − (45)

The fraction of the bubbles in the dense zone is:

0 gbb

b

uf

υ= − (46)

The absolute bubble rise velocity bυ is calculated as a function the isolated bubble rise velocity: ( )0.5

:RG0 0.711 b g ge bgd u u υ= + − − (47) The bubble diameter is estimated by the Wen-Mori correlation [4]:

( )

( ) ( )

1:RG

1 0:RG

0.30 expD

D D

lD D b b

l lD b b

L l d dD d d

=

= =

⎛ ⎞ −= − −⎜ ⎟

−⎝ ⎠ (48)

The initial bubble diameter and the maximum bubble diameter are estimated by eqs.(26) and (27), respectively:

( )( ) ( )0.4

0 0:RG0.2

1.38 101000

D Dl lg mb bu u d

g= =⎛ ⎞= − −⎜ ⎟

⎝ ⎠ (49)

( ) ( )( )( )( )( )22.7

0.41 1

:RG :RGmin 0.652 , 2p

D D

dtl l

b D g mb

ud A u u

g= =

⎡ ⎤⎢ ⎥

= −⎢ ⎥⎢ ⎥⎣ ⎦

(50)

The emulsion to freeboard elutriation rate *iK of a fraction of particles with average

diameter dpi is evaluated by the Zenz and Weil correlation [19]. The total entrainment rate *tK

is then obtained by adding the rates of each respective fraction of particles. The catalyst density in the freeboard is a function of the gas-solids slip velocity, which is calculated on the basis of the correlation of Patience et al. [20], as shown in eq.(51):

0.4101 5.6 / 0.47

gfsf

gf t

uu

Fr Fr= −

+ + (51)

The freeboard voidage is then calculated by eq.(52):

*

0 1 tf

p sf

Ku

ερ

= − − (52)

The pressure drop throughout the regenerator is calculated from the solids static head as shown in eq.(53): ( ) ( )RG :RG :RG1 1p e e D p f FP f gL gLρ ε ρ ε∆ = − + − (53)

Simulation of Stripper and Slide Valves

The pressure drop throughout the stripper is calculated from the solids static head as shown in eq.(54): ( )ST :ST1p mf DP gLρ ε∆ = − (54)

Page 19: Model Predictive Control of a Fluid Catalytic Cracking

The catalyst circulation rate at the entrance and exit of the regenerator was correlated with the slide valves opening of the stripper and regenerator standpipes and the pressure drop by eq.(55) [21]:

( )( )

0.52 2 0.50 :SVSV 2 2

0 :SV

2 1tC p SV

t

A AW k PA A

ρ ε⎛ ⎞

= − ∆⎜ ⎟−⎝ ⎠&

(55)

APPENDIX III: Nomenclature

A cross-sectional area (m2) xcy catalytic coke yield on feed basis (wt%)

cy total coke yield on feed basis (wt%) D diameter (m) dp catalyst particle mean diameter (m3) cp specific heat (kcal mol-1K-1) Ex, Ec activation energy of reaction to x, c (kcal mol-1) fb, fe bubble, emulsion phase volume fraction Fib molar rate in bubble (mol s-1) cie molar concentration in emulsion (mol s-1) Fif molar rate in freeboard (mol s-1) Ht bubble-emulsion heat interchange (kcal m-3 s-1) KH heat interchange rate group (kcal m-3 s-1) KMi mass interchange rate group (mol m-3 s-1) KRjb reaction rate group of reaction j - bubble phase (mol m-3 s-1) KRje reaction rate group of reaction j - emulsion phase (mol m-3 s-1) KRjf reaction rate group of reaction j - freeboard (mol m-3 s-1) KSV1, KSV2 characteristic constant of slide valve SV1, SV2 Kti bubble-emulsion gas interchange coefficient (s-1) kx, kc pre-exponential factor of reaction to x or c L height (m) l dimensionless height MWF feed molecular weight nx, nc catalyst decay exponent of reaction to x or c Qb enthalpy rate in bubble phase (kcal s-1) QC enthalpy rate of catalyst (kcal s-1) Qge enthalpy rate of gas in emulsion phase (kcal s-1) Qloss heat loss from the dense bed (kcal s-1) P pressure (Pa) T temperature (°C) TRX riser reactor temperature (°C) tC catalyst residence time (s) tdead time lug in standpipe or liftline (s) ug superficial gas velocity (m s-1) ut catalyst particle terminal velocity (m s-1) V volume (m3) WHSV weight hourly space velocity (hr-1)

CW& catalyst circulation rate (kg s-1)

Page 20: Model Predictive Control of a Fluid Catalytic Cracking

FW& gas-oil feed rate (kg s-1) yx gas-oil conversion on feed basis (wt%) ∆Hcrack heat of catalytic cracking (kcal kg-1) ∆HRj heat of reaction j (kcal mol-1 )

Greek Letters αij stoichiometric coefficient of component i in reaction j εb voidage of bulk catalyst εmf voidage at minimum fluidization εe dense bed emulsion void fraction εf freeboard void fraction εr riser void fraction ρp catalyst density (kg m-3)

Subscripts g gas phase s solids phase b bubble phase e emulsion phase f dilute phase

Unit Section Subscripts D dense phase or bottom section C cone intermediate section F dilute phase or top section CY regenerator cyclone RS riser RG regenerator ST stripper SP standpipe LL liftline