Вы находитесь на странице: 1из 19
Modeling || the Dynamic and Static Behavior of Chemical Processes What goes on in the modeller’ head is not purely formalizable, either in abstract terms....or in taxonomic views....It has structure, it has technique that can be taught and learned, but involves also a personal touch, not only in trivialities but in deeper considerations of skill and suitability. R.Aris and M. Penn* In order to analyze the behavior of a chemical process and to answer some of the questions raised in previous chapters about its control, we need a mathematical representation of the physical and chemical phenomena taking place in it. Such a mathematical repre- sentation constitutes the mode/ of the system, while the activities leading to the construction of the model will be referred to as model- ing Modeling a chemical process is a very synthetic activity, requiring the use of all the basic principles of chemical engineering science, such as thermodynamics, kinetics, transport phenomena, etc. For the design of controllers for chemical processes, modeling is a very criti- cal step. It should be approached with care and thoughtfulness. The purpose of the following two chapters is: 1. To explain why we need to develop a mathematical description (model) of a chemical process as a prerequisite to the design of its controller ‘The Mere Notion of a Mathematical Model,” Int. J. Math, Modeling, 1, | (1980). 44 Modeling the Dynamic and State Behavior af Chemical Processes Part 2. To describe a methodology for the modeling of a chemical proc- ess using the balance equations and provide examples of its implementation 3. To determine the scope and the difficulties of the mathematical modeling for process control purposes It should be noted that the subsequent chapters do not constitute a complete treatment of all the aspects on mathematical modeling, but it is limited to those of interest for process control. Development 4 of a Mathematical Model Consider a general processing system with its associated variables as shown in Figure 2.1. To investigate how the behavior of a chemical process (i.e., its outputs) changes with time under the influence of changes in the external disturbances and manipulated variables and consequently design an appropriate controller, we can use two different approaches: 1, Experimental approach: In this case the physical equipment(s) of the chemical process is available to us. Consequently, we change deliberately the values of various inputs (disturbances, manipulated variables) and through appropriate measuring dev- ices we observe how the outputs (temperatures, pressures, flow rates, concentrations) of the chemical process change with time. Such a procedure is time and effort consuming and it is usually quite costly because a large number of such experiments must be performed. 2. Theoretical approach: It is quite often the case that we have to design the control system for a chemical process before the proc- ess has heen constructed. In such a case we cannot rely on the experimental procedure, and we need a different representation of ‘the chemical process in order to study its dynamic behavior. This, representation is usually given in terms of a set of mathematical equations (differential, algebraic) whose solution yields the dynamic or static behavior of the chemical process we examine. 46 “Modeling the Dynamic and Static Behavior af Chiwrical Processes Part In this text we discuss both approaches for the development of a model for a chemical process. Initially, we will examine the theoretical approach, leaving the experimental for subseayént chapters (Chapters 16 and 31). 4.1 Why Do We Need Mathematical Modeling for Process Control? Let us repeat that our goal is to develop a control system for a chemical process which will guarantee that the operational objectives of our process are satisfied in the presence of ever-changing disturbances. Then, why do we need to develop a mathematical description (model) for the process we want to control?” In the introductory paragraphs earlier we noted that often the physi- cal equipment of the chemical process we want {8 control have not been constructed. Consequently, we cannot experiment to determine how the process reacts to various inputs and therefore we cannot design the appropriate control system. But even if the process equipment is availa- ble for experimentation, the procedure is usvally very costly. Therefore, we need a simple description of how the prdvess reacts to various inputs, and this is what the mathematical siaqels can provide to the control designer. Let us demonstrate now in terms of some exmples the need for the development of a mathematical model before we design the control system for a chemical process. Example 4.1: Design an Integral Contraller for a Stirred Tank Heater Consider the problem of controlling the temperature of a liquid in a tank using integral control (Example 2.12) From Figure .8 we notice that the quality of the control depends on the value of the parameter a’, But ‘the question is: How does a’ affect the quality of ¢ontrol, and what is its best value? To answer this question we need to kndw how the value ofthe liquid temperature Tis affected by changes in the value of the inlet temperature 7; or the integral action of the con(raller. This is given by eq. (2.7), which constitutes the mathematical mode! af the tank with integral control Example 4.2: Design a Feedforward Contralles for a Process In the feedforward control arrangement shaw in Figure 4.1 we mea sure the value of the disturbance and we anticinal what its effect will be ‘on the output of the process that we want to ont"p1, In order to keep the Chap. 4 Development of a Mathematical Model 47 Disturbance Contoter Proce Manipusted ‘Oupar varisble Figure 4.1 Feedforward control configuration, value of this output at the desired level, we need to change the value of the manipulated variable by such an amount as to eliminate the impact that the disturbance would have on the output. The question is: By how much should we change the manipulated variable in order to cancel the effect of the disturbance? To answer this question we must know the following two relationships: output « fi(disturbance) output = f;(manipulated variable) which are provided by a mathematical model of the process. Indeed, if ‘the output is to remain the same, the manipulated variable must take such a value that ‘fildisturbance) ~ f,{manipulated variable) = 0 This example demonstrates very vividly how important mathematical modeling is for the design of a feedforward control system. In fact, with- ‘out good and accurate mathematical modeling we cannot design efficient feedforward control systems. Example 4.3: Design of an Inferential Control System In the inferential control scheme shown in Figure 4.2 we measure the ‘measured output and try to regulate the value of the unmeasured control objective at a desired value, Since the control objective is not measured directly, it can only be estimated from the value of the measured output if a relationship such as the following is available: control objective = flmeasured output) Such a relationship in turn is not possible if we do not have a mathemati- cal representation of the process (mathematical model). Once the value of the control objective can be estimated from a relationship such as the above, it can be compared to the desired value (set point) and the control- ler can be activated for appropriate action as in feedback control. ‘We notice, therefore, that the availability of a good mathematical model for the process is indispensable for the design of good inferential control systems. 48 Modeling the Dynamic and State Behawor of Chemcal Processes Part I Manipuated Controted tanable Measured Contaler ome Figure 4.2 Inferential control configuration, The three examples above indicate very clearly that mathematical modeling of a process is of paramount importance for the design of good and efficient control systems for a chemical process. In the follow- ing sections we develop a methodology for the concise modeling of chemical processes. 4.2 State Variables and State Equations for a Chemical Process In order to characterize a processing system (tank heater, batch reactor, distillation column, heat exchanger, etc.) and its behavior we need: 1. A set of fundamental dependent quantities whose values will describe the natural state of a given system 2. A set of equations in the variables above which will describe how the natural state of the given system changes with time For most of the processing systems of interest to a chemical engineer there are only three such fundamental quantities: mass, energy, and momentum. Quite often, though, the fundamental dependent variables cannot be measured directly and conveniently. In such cases we select, other variables which can be measured conveniently, and when grouped appropriately they determine the value of the fundamental variables. Thus mass, energy, and momentum can be characterized by variables such as density, concentration, temperature, pressure, and flow rate. These characterizing variables are called state variables and their values define the state of a processing system, ‘The equations that relate the state variables (dependent variables) to the various independent variables are derived from application of the conservation principle on the fundamental quantities and are called state equations. Chap. 4 Development of a Mathematica! Mode! 49 The principle of conservation of a quantity S states that: [ene ion ors] [, flow of 5 ] [, flow of S ] within a system | _Lin the system] [out of the system, time period ‘time period time period an amount of S] amount of S generated vi consumed within the system the system time period time period The quantity S can be any of the following fundamental quantities: Total mass Mass of individual components Total energy Momentum Remark, It should be remembered that for the physical and chemical Processes we will be studying, the total mass and total energy cannot be generated from nothing; neither do they disappear. Let us review now the forms used most often for the balance equa- tions. Consider the system shown in Figure 4.3. We have: Boundary defining se system 2 outta Figure 4.3 A general system and its interactions with the external world, 50 Modeling the Dynamic and State Behawor of Chemical Processes Part Total mass balance: (pV) ao iF; — nF (4.1 ae i Si (a) ‘Mass balance on component A: dian) _dlea¥) HW yok 5 eye 4. at at NEI Zi ewtierY (Aub) Total energy balance: aE _d(U+K+P)_ rai ai Rhu Zeb se Qe W, (Ale) The variables appearing in the equations above have the following meaning: density of the material in the system density of the material in the th inlet stream jensity of the material in the jth outlet stream jotal volume of the system /olumetric flow rate of the ith inlet stream F, = volumetric flow rate of the jth outlet stream ny = number of moles of component A in the system ¢a= molar concentration (moles/volume) of A in the system ¢a,= molar concentration of A in the ith inlet a; = molar concentration of A in the jth outlet r= reaction rate per unit volume for component A in the system +h, = specific enthalpy of the material in the ith inlet stream +h = specific enthalpy of the material in the jth outlet stream U, K, P=imernal, kinetic, and potential energies of the system, respectively Q-amount of heat exchanged between the system and its surroundings per unit time W,= shaft work exchanged between the system and its sur- roundings per unit time By convention, a quantity is considered positive if it flows in the system and negative if it flows out. The state equations with the associated state variables constitute the ‘mathematical model of a process, which yields the dynamic or static behavior of the process. The application of the conservation principle Chap. 4 Development of a Mathemavcal Mode! 61 as defined by eqs. (4.1) will yield a set of differential equations with the fundamental quantities as the dependent variables and time as the independent variable. The solution of the differential equations will determine how the fundamental quantities, or equivalently, the state variables, change with time; that is, it will determine the dynamic behavior of the process. If the state variables do not change with time, we say that the process is at steady state. In this case, the rate of accumulation of a fundamental quantity Sper unit of time is zero, and the resulting balances yield a set of algebraic equations. Example 4.4: State Variables and State Equations for a Stirred Tank Heater; Its Static and Dynamic Behavior Consider the stirred tank heater of Example 1.1 (Figure 1.1), The fundamental quantities whose values provide the information about the heater are: (a) The total mass of the liquid in the tank (b) The total energy of the material in the tank (6) Its momentum. ‘The momentum of the heater remains constant even when the distur- ‘ances change value and will not be considered further. Let us now identify the state variables for the tank heater. Total mass in the tank: total mass = pV = p4h (42) where p the density of liquid, V the volume of liquid, A the cross- sectional area of the tank, and / the height of the liquid level Total energy of the liquid in the tank: E=U+K+P but since the tank does not move, dK/dt = dP/dt = O.and dE/dt = dU/dt. For liquid systems, wu att a at where #7 isthe total enthalpy ofthe liquid inthe tank. Furthermore, H = pVof{T~ Tea) pAbe T= Ta) (43) where cj = heat capacity ofthe liquid in the tank ‘Tera reference temperature where the specific enthalpy of the ie uid is assumed to be zero. From eqs. (4.2) and (4.3) we conclude that the state variables for the stirred tank heater are the following; 52 ‘Modeling tne Dynamic and Static Behavior of Chemical Processes Part State variables: A and T while the Constant parameters: p, A, 6, Tw are characteristic of the tank system, Note. It has been assumed that the density p is independent of the temperature. Let us now proceed to develop the state equations for the stirred tank heater. We will apply the conservation principle on the two fundamental quantities: the total mass and the total energy. Total mass balance: [scumuistion “4 [ input of ] ce | total mass J _ [total mass|_ [total mass. time time time or (4) where F, and F are the volumetric flow rates [i.c., volume per unit of time (ft/min, or m/min) for the inlet and outlet streams, respectively ‘Assuming constant density (independent of temperature), eq. (4.4) becomes Atapor (4a) dt Total energy balance: [esamusionet] [izeatet } [onto | totalenergy _|_Ltotalenergy.| [total energy. time time time lee aaa ‘by steam time or dlpah oT = Tx} dt ‘where Q is the amount of heat supplied by the steam per unit of time. The equation above can take the following simpler form (assume that Tram 0}: = pF as T,~ Tn) ~ PFT = Tw) + Q (4.5) in Fr+ 2 Pe (45a) ah) or Chap. 4 Development of a Mathematical Mode! 53 Additional oe ‘manipulations on eq. (4.5a) yield AAT) 4 ya ar var the an Ee TR at at or wa rer 8 on me Summarizing the modeling steps above, we have: State equations: (44a) Fa F(T, 1) 2 (430) Pe “The variables in eqs. (44a) and (4.50) canbe classified as follows (see also Section 2.1): State variables: h, T Output variables: h, T (both measured) Input variables Disturbances: T;, F, Manipulated variables: Q, F (for feedback control) F (for feedforward control) Parameters: A, p, Cy The state equations (4.4a) and (4.5b), with the state variables, the inputs, and the parameters, constitute the mathematical model of the stirred tank heater. We need only solve them in order to find the tank's dynamic or steady-state behavior. Let us now study the dynamic and static behavior of the stirred tank heater using the state equations (4.4a) and (4,56), We will assume that initially the tank heater is at steady state (ie, nothing is changing). This situation is described by the state equations if the rate of accumulation {left-hand sides of (4.4a) and (4.5b)] is set to zero: Fis Fin 0 F,(Tig~ T+ B The subscript s denotes the steady-state value of the corresponding varia- ble ‘The system wil be disturbed from the steady-state situation if any of the input variables changes value. Let us examine the following two situations: 1. Consider that the inlet temperature 1’, decreases by 10% from its steady-state value. The liquid level will remain the same at the 54 “Modeling the Dynamic and Static Behavior of Chemical Processes Part I Od stcady state "New steady state Figure 4.4 Temperature response of a stirred heater to a step decrease in inlet temperature steady-state value h,, since T; does not influence the total mass in the tank [see also eq. (4.4a)]. On the contrary, the temperature of the liquid will start decreasing with time. How the temperature T cchanges with time will be determined from the solution of eq. (4.50) using as initial condition the steady-state value of T: TU =0)= 7, Figure 4.4 indicates the static and dynamic behavior of the tank for this case. We observe that after a certain timie the tank heater again reaches steady-state conditions. 2. Consider that initially the tank heater is at steady state. Then, at time ¢ = 0, the inlet flow rate decreases by 10%. It is clear that both the level and the temperature of the liquid in the tank will start changing [notice that F, is present in both state equations (4.4a) and (4.5b)]. How A and T change with time will be given from the solution of eqs. (4.4a) and (4.5b) using as initial conditions h(t=0)=h, and T(t =0)= 7, Figure 4.5 summarizes the static and dynamic behavior of the tank heater for this case. ld steady state New steady state "Nei iad sate ° Time ° Time Figure 4.5 Dynamic response of a stirred tank heater to a step decrease in inlet flow rate. Chap. 4 Development of a Mathematical Mode! 55 Remark. tis worth noting that after F; has changed, the level h and the temperature 7’ reach their new steady states with different speeds. In particular, the level, A, achieves its new steady state faster than the temperature. In a subsequent chapter we will analyze the reasons for such behavior. 4.3 Additional Elements of the Mathematical Models In addition to the balance equations, we need other relationships to express thermodynamic equilibria, reaction rates, transport rates for heat, mass, momentum, and so on. Such additional relationships needed to complete the mathematical modeling of various chemical and/or physical processes can be classified as follows: ‘Transport rate equations They are needed to describe the rate of mass, energy, and momen- tum transfer between a system and its surroundings. These equations are developed in courses on transport phenomena. Example 4.5 The amount of heat Q supplied by the steam to the liquid of the tank heater (Example 4.4) is given by the following heat transfer rate equation: Q=UA(Ts-T) where U = overall heat transfer coefficient y= total area of heat transfer T.= temperature of the steam. Kinetic rate equations They are needed to describe the rates of chemical reactions taking place in a system. Such equations are developed in a course on chemical Kinetics. Example 4.6 The reaction rate of a first-order reaction taking place in a CSTR is given by rm keen Modetng the Dynamic and Static Behavior of Chemical Processes Part I durhere ko shir E R Teen preexponential kinetic constant activation energy for the reaction ideal gas constant temperature and concentration of the reacting fluid. Reaction and phase equilibria relationships HThese are needed to describe the equilibrium situations reached Jing a chemical reaction or by two or more phases. These relation- ‘ps are developed in courses on thermodynamics. Fxample 4.7 Consider a liquid stream composed of two components A and B at a high pressure py and temperature 7). If the pressure py is larger than the pubble-point pressure of the liquid at temperature 7), no vapor phase will be present. The liquid stream passes through a restriction (valve) and is “flashed” in a drum; that is, its pressure is reduced from py to p (Figure 4.6). This abrupt expansion takes place under constant enthalpy. If the pressure p in the drum is smaller than the bubble-point pressure of the ‘quid stream at the temperature 7;, the liquid will partially vaporize and |wo phases at equilibrium with each other will be present in the flash Arum. "The thermodynamic equilibrium between the vapor and liquid phases imposes certain restrictions on the state variables of the system, and ‘must be included in the mathematical model of the flash drum if it is to be consistent and correct. These equilibrium relationships, as known from chemical thermodynamics, are: 1. Temperature of liquid phase = temperature of vapor phase v a Figure 4.6 Flash drum unit Chap. 4 Development of a Mathematical Mode! 87 2. Pressure of liquid phase = pressure of vapor phase 3. Chemical potential of component / in the liquid phase = chemical potential of component i in the vapor phase ‘The equilibrium relationships introduce additional equations among the state variables of a system. Care must be exercised so that all the equilibrium relationships are accounted for. Equations of state Equations of state are needed to describe the relationship among the intensive variables describing the thermodynamic state of a system. ‘The ideal gas law and the van der Waals equation are two typical ‘equations of state for gaseous systems. Example 4.8 Let us return to the flash drum system discussed above in Example 4.7. For the vapor phase, from the ideal gas law, we have PV apa = (moles of A + moles of B): RT 46) But mass of A + mass of B moles of A + moles of B = — a average molecular weight Therefore, from eq. (4.6) we have pane MESSOFA mass FBP “ Vase RT {average molecular weight) Considering that average molecular weight = yaMa+ YeMe we have Posse = Pa tyaM a+ YoMa] (46a) RT where ya and ys are the molar fractions of components A and Band Mx and Mg are the molecular weights of A and B. Equation (4.6a) indicates a relationship among the state variables of the flash drum and must be included in the mathematical mode! of the flash drum. ‘Starting from an appropriate equation of thermodynamic state for the liquid phase of the flash drum, we can develop an expression for its density of the form Prunus = (Ts Xn) where 2¢4 is the molar fraction of component A in the liquid. 68 Modeling the Dynamic and Static Behavior of Chemical Processes Part ‘chap. 4 Development of a Mathematical Model 59 ear that the temperature of the outlet, Tw, will remain the same until 4.4 Dead Time thechange reaches the end ofthe pie Then we wl observe the temper In all of the modeling examples discussed in earlier sections it was that ie change of he ute iemperatar lows tea prego he assumed that whenever a change takes place in one of the input vari- change of the inlet temperature with a delay of ty seconds. tis the dead ables (disturbances, manipulated variables), its effect is instantaneously time and from physical considerations itis easy to see that ‘observed in the state variables and the outputs. Thus whenever the feed mane AP Composition, ca, or the feed temperature, 7), or the coolant tempera ty = Yolume of the pipe | _4-L_ a= seconds ture, T.,, change in the CSTR of Figure 1.7, the effect of the change is Volumetric flow rate A:Uy ~ Un felt immediately and the temperature T or concentration cx of the where Us. is the average velocity of the fluid over the cross-sectional area, outlet stream start changing. of the pipe, assumed to be constant. Functionally, we can relate Ti, and ‘The oversimplified picture given above is contrary to our physical Tu a8 follows experience, which dictates that whenever an input variable ofa sytem Todt) Tet) an ime interval (short or long) during which no effec changes, there time inte oor rate eene interval is called dead The dead time is an important clement inthe mathematical modeling ‘observed on the outputs of the system. This t ocity of chemical processes and has a serious impact on the design of effective time, or transportation lag, or pure delay, or distance-velocity lag. controllers. AS we will see in Chapter 19, the presence of dead time can very easily destabilize the dynamic behavior ofa controlled system. Example 4.9 sider the fw of an incompressible, onreating liquid through a pins Figure 47) Tf he pipe completely thermally insulated andthe at generated by the fcion ofthe wing Mulds nelle, itis easy 1 eee Ceiba at steady tte the temperature Tx ofthe outlet steam wil be 0 SGealto tha ofthe inet, Te, Assume now that starting at =O, the Modeling ‘Guperaute ofthe let changes as shown by curve in Figure 7b. tis In this section we examine some typical chemical processes and develop their mathematical models. Example 4.10: Mathematical Model of a Continuous Stirred Tank Reactor (CSTR) Consider the continuous stirred tank reactor system discussed in Example 1.2 (Figure 1.7). A simple exothermic reaction A ~ B takes place in the reactor, which is in turn cooled by a coolant that flows through a Jacket around the reactor. ‘The fundamental dependent quantities for the reactor are: (@) Total mass of the reacting mixture in tank (b) Mass of chemical A in the reacting mixture (©) Total energy of the reacting mixture in the tank Remarks 70 Tine 1. The mass of component B can be found from the total mass and the © ‘mass of component A. Therefore, it is not an independent funda- ‘mental quantity. Figure 47 (a) Pipe flow of Example 4.9; (0) delayed response of exit 2. The momentum of the CSTR does not change under any operating temperature to inlet temperature change. conditions for the reactor and will be neglected. 60 Modelng the Dynamic end State Behavior of Chemical Processes Part Let us apply the conservation principle on the three fundamental quantities: Total mass balance: f tcrumulaon] [smc] [ouster] [lostmaesenentd Loftotal mass] [total mass _Ltotal mass} ,|__or consumed. - time time time time 40) pF pF £0 48) ai where p.p = densities of the inlet and outlet streams. , ‘volumetric flow rates of the inlet and outlet streams, ft/min or m/min volume of the reacting mixture in the tank ‘Mass balance on component A. [accumulation] [input] [output] [shareware 4] ora | bora] Lora | | duetoreaction “time time time time (ns) _ dlea¥) a at where r= rate of reaction per unit volume CagCa = molar concentrations (moles/volume) of A in the inlet and outlet streams and a= number of moles of A in the reacting mixture en Fi~ ea -1V oo) Total energy balance: accumulation of] (“se time input of total] [ output of wal cre removed [ nergy with feed.) [energy with outlet) | by coolant time time time In the balance above we have neglected the shaft work done by the impeller of the stirring mechanism. The total energy ofthe reacting mminture fe B= U+K+P where U is the internal energy, K the kinetic energy, and ie olen Tal energy ofthe reacting minut, Therefore, assuming that he reac tor does not move (ie, dK/d = dP/ai~ 0), the lechand side ofthe total energy balance yields Chap. 4 Development of a Mathematica! Model 61 dE _d(U+K+P)_ du dt dt at Since the system is a liquid system, we can make the following approxi- matior [ accumulation of total [ accumulation of total") energy of the du. att | emihalpy ofthe material in the CSTR |=“ *“q ~| material in the CSTR pet unit time per unit time Furthermore, input of total energy with feed per unit time = p,F\h(Ti) and output of total energy with the outlet stream per unit time = pF A(T) where hi is the specific enthalpy (enthalpy per unit mass) of the feed stream and h is the specific enthalpy of the outlet stream. Consequently, the total energy balance leads to the equation 1. pEn(T)~pER(T) 0 (4.10) where Q is the amount of heat removed by the coolant per unit time. Equations (4.8), (4.9) and (4.10) are not in their final and most conve- rient form for process control design studies. To bring them to such form we need to identify the appropriate state variables. Characterize Total Mass, We need the density of the reacting mixture, p, and its volume, V. The density will be a function of the concentration cx and cy and of the temperature T. Quite often the dependence of p on cs, ¢a, and T is weak and the density can be considered constant as the reaction proceeds. Therefore, the left-hand side of eq. (4.8) yields ae white p, Under the assumption above, Vis the only state variable that is needed to characterize the total mass. Then eq. (4.8) becomes av. oon-F 4 "7 (48a) Characterize the Mass of Component A. This is simple. From eq. (4.9) we realize that the state variables needed are c, and V. Aigebraic manipula- tions on eq. (4.9) lead to UeAV)_ AV, den nt VEN cu caF ~ koe MY a aT ae“ : 62 Modeling the Dynamic and Static Behavior of Chemical Processes Part! ae : a Vv She eR F + ¢a,Fi~ CaF - koe ®™ caV dt CnC ) and finaly, dey _Fi -! den Ficey,— cn) = hoe Cn co) dt 7 . y J [Note. We have introduced r = koe **7Cu.] Characterize the Total Energy. We know from thermodynamics that the enthalpy of a liquid system is a function of the temperature and its composition: He H(T, nam) where na and ng are the moles of A and B in the CSTR. Differentiating the expression above, we take OH day , aH day (aay i dt” dow dt But na oH ang = AMT cp Haag Fn RlT) where c, is the specific heat capacity of the reacting mixture and Fixand Fis are the partial molar enthalpies of A and B, Furthermore, from eq. 49), drin ACV) «oF, caF = a dt and a similar balance on component B, Substitute the quantities above in ea. (4.11) and take BH yy, TU + Alen, Fi~ ea — V1 + Fial-coF + V1 dt dt Substitute dH /dt by its equal from the total energy balance (69. (4.10)] and take ar ve, ar (4.108) = -Bislen Fi = eaF - rV)~ Bial-coF + rV)+ pF sh - PPh ~ Chap. 4 Development of a Mathematical Model 63 Let us now notice that FpihdTi) = FipshdT) + pie p(T s ~ T)) = File AAT) + pie - TY) and Foh(T) = FlenHA(T) + ¢sfo(T)] Consequently, eq. (4.108) becomes Pi ey En ypc reek + Alar 4 KEE — ftarv + Exei(Hn+ Fipic)(Ti~ T)- Pky - 5a - or ar of PVey Ge = FipscglTi~ T)+ (Ha flay ~ Finally, since (i, - As) =(-AH,) and p= pis Cp = Crp eat of reaction at temperature T, ar AH) Fe pyr, 1A 2 (4.100) Pree From eq, (4.10b) we conclude that temperature 7 is the state variable characterizes the total energy of the system. able that Summarisng al ihe spe above late mathematical modeling ofa CSTR, we have the following: sting of State variables: V; ¢x,T State equations: wv a dex av aT Fy sun, Geo Tn Dt Shae en where J = (-AH)/pcp. F (48a) (cay en) = koe Fen (49a) (4.10b) Output variables: V, cx, T Input variables: cn, F, Ti, QF (when feedback control is used) Among the input variables the most common disturbances are: Disturbances: Cn Fi Ts while the usual manipulated variables are: Manjpulated variables: Q, F (occasionally F, or T:) 64 Modeling the Dynarnec and Static Behavior of Chemical Processes Part The remaining variables are parameters characteristic of the reactor sys- tem: Constant parameters: p, Cy (~AH,), ko, E (activation energy), R In the presence of changes in the input variables, the state variables change. Integration of eqs. (4.8a), (4.9a), and (4.10b) yields (1), e4(t), and T(1) as functions of time. The steady-state behavior of the CSTR is given by eqs. (4.8a),(4.9a), ‘and (4.10b) if their left-hand sides are set equal to zero. Example 4.11: Mathematical Model of a Mixing Process ‘Two streams | and 2 are being mixed in a well-stirred tank, producing ‘a product stream 3 (Figure 4.8). Each of the two feed streams is composed of two components, A and B, with molar concentrations ca,, Cx, and Cas, ap Fespectively. Also let Fi and F2 be the volumetric flow rates of the two streams (ft/min, m’/min) and T; and T; their corresponding tem- peratures. Finally, let ¢4,, Cay, Fs, and 7 be the concentrations, flow rate, and temperature of the product stream. A coil is also immersed in the liquid of the tank and it is used to supply heat to the system with steam, or remove heat with cooling water. The fundamental quantities needed to describe the mixing process are: (a) Total mass in the tank (b) Amounts of components A and B in the tank (©) Total energy (4) Momentum of the material in the tank Remarks 1. The momentum does not change under any operating conditions and it will be neglected in further treatment, 2, We only need to consider two of the following three quantities: total mass, mass of A, mass of B, The third can be computed from the other two, Sweam 1 Stream 2 ante cay TF ep eiteat ates Seremoved) seam 3 ay Tes Figure 4.8 Mixing process Chap.4 Development of @ Mathematical Model 65 Consider now the balances on the fundamental quantities: Total mass balance: seem of ua] [meat output of total ‘mass in the tank _|_[mass in the tank mass from the time time time dov) dt where pi, pa, and py are the densities of streams 1, 2, and 3, respectively. Since the content of the tank is well mixed, the density of the product stream ps is equal to the density of the material in the tank, p (i. py = 9) V is the volume of the material in the tank and is equal to the product of the cross-sectional area of the tank, A, and the height, h, of the liquid level: (AF i+ LF) - pF (4.2) Vash In general, the densities, p, p, and po depend on the corresponding con- ‘centrations and temperatures: P= Ps=Klery Coy Ts) Pr=fleay cry Tr) — pr=Sleay Coy T2) Usually (but not always) the dependencies above are weak and we assume that the densities are independent of the concentrations and tempera tures. Therefore, we assume that P= p=ps=p This transforms eq. (4.12) to the following: av _ a aA BaF + FFs 2 GAG Fit Fa F (4.128) Balance on component A: sccumsstiono?) [totlipat of) [ot outer compenentAin || component come | thetank _}_L inthetank } | from the tank | time time time or dca¥) et) (Ca) Fi + CaF) — Cass a ) = ay den... v VE eM eu F + em Fi) CasFs Tet ON Gp Men emP Dd ~ CasF (4.13) Substituting d¥/dt by its equal from eq. (4.12a), we take 66 Modeling the Dynamic and State Behavior of Chemical Processes Pati Ves ese F)~BlmlenPr + euho)~enFs and since cx = ¢4, due to the well-stirred assumption, V8 (64, ea)Fi Hen ~ Cuda -EQges (4.13a) Total energy balance: ‘accumulation of] input of total energy four of total energy’ ] {_totalenergy J _ | with feed streams time time [heat added or removed) L___ with the coit time The total energy of the material in the tank is E =U (internal) + K (kinetic) + P (potential) Since the tank is not moving, dK/dt = dP/dt = 0. Thus dE/dt = dU/dt and for liquid systems, wan ad were H isthe total enthalpy ofthe material inthe tank, Furthermore, input of total energy ( with ed seam ) cath per unit time and output of total energy with product stream} = pF shy per unit time where fy, ha, and hy are the specific enthalpies (enthalpy per unit mass) of streams |, 2, and 3. Due to the perfect stirring assumption, the specific enthalpy of the material in stream 3 is the same as the specific enthalpy of the material in the tank. Thus Mah, Consequently, the total energy balance yields OED «OCF hy + Fite) ~pF vhs 2 44) Chap. 4 Development of a Mathematical Model 67 The question now is how to characterize hs, hz, and h; in terms of other variables (i., temperatures, concentrations, etc.). We know that AATs) = hfTe) + ¢9,(Ts~ Te) (4.15) I(T) = hi(To) + ep (Ts ~ To) (4.156) AAT2)= hATo) + Cp T2~ To) (4.150) ‘where Ty is the reference temperature. At this temperature PhSTo) = Carll + Coll + cas S5(To) (4.16a) phi(Ta) = CayFn + Cftn+ ca, Aso) (4.166) Phy To) = Casa + CoH + Cx AHs,(To) (4.160) where #4 and Hs are the molar enthalpies (enthalpy per mole) of com- ponents and Bat temperature Ts. Afls, AH, and Alls, are the heats of solution for streams 1, 2, and 3 per'mole of A at temperature T', Substituting eqs. (4.15a,b,¢) and (4.16a,b.c) into the total energy balance ca. (4.14), we take AV Cayba + Coyle + Cry Alls) + pV CoA Ts ~ To)) dt = Filcall n+ cna + ca, Mffs,) + pF icy(T1 - To) + Fuca a + CogAn + Ca; DAs) + pF 2ep(T2 ~ To) ~ Fa(caslla + cnylls + cay Mffsy) ~ pF scp Ts - To) = Q (balance on A) AUC TO) , | eR, Fit ens] edt ana es a + +. d(Ven) [Percocet Med ‘= 0 (balance on B) = Fica, AAs, + pF cp (Ts ~ To) + Fxtxy Mls, + PF x¢p,(T2 ~ To) ~ Fea AAs, ~ pFacy(Ts ~ To) t Q or ay a oan “at Peo Ts ~ To) + MA sila Fi + CagF2 ~ CayF a] Pen FP Bea(Ts~ To) c , = Fin, Mis, + pF icy (Ty ~ To) + Fxtay Mf sy + PF 1¢p,(T2 - To) = Fea, Mls, - pF ¢,(Ts~ To) £ 68 Modeling the Dynamic and Statie Behavior of Chemical Processes Part and finally, pen¥ a = cuFlQAs— Bis] + emg Fldits, - 4715) it + PF ilep4(Ts ~ To) ~ ¢o(Ts~ Tol + pF ep T2— To) ~ ¢o(To- Tol] * Q If'we assume that Cy, = Cy, = Cyy = Con WE have: a7 va pew a eu Filly, SAg) + cm FAAMs, ~ Aff] 4 pFie( Ti 13) +0F ich T:~T)£Q (Alda) ‘Summarizing the steps above, we have: Sate variables: Vey Ts ‘Sate equations: Wes Fa-Fs (4.128) at V E21 «(on,— Cu) Fi + (Ena ea)F (4.138) i pep To en FMA, AA} + cogFilSfls~ Aft) i + pFyey(Ti~ Ts) + pF 209(T2- 11) # (ata) Input variables: Fs, ¢ny. Ti, Fs Cnn To, Fs (for feedback control) Output variables: V (or equivalently the height of liquid level, h), Cay Ts Parameters (constant): p, Cr, Ms, AA1sy, AAs, Remarks 3. Usually a mixing tank is equipped witha cooling or heating coil or jacket through which flows a coolant (if heat is released during the mixing of the two solutions) or a heating medium (if heat is absorbed during mixing) in an attempt to Keep the mixing isother- 4. Hike heats of solution ae strong functions of concentration (ie. if [Afs,~ Aff) and [AHs, AM) are not small quantities), then from the total energy balance, eq (4.14a), we notice that tempera- ture T; depends strongly on the concentrations of the fed streams fand their temperatures. If on the other hand, {Affs,- Als) and [a7%5,_ Affs are nearly zero, then 7 depends basically only on Tyand T Cchap.4 Development of a Mathematical Model 69 Example 4.12: Mathematical Model of a Tubular Heat Exchanger Consider the shell-and-tube heat exchanger shown in Figure 4.9. A liquid flows through the inner tube and it is being heated by steam that flows countercurrently around the tube. The temperature of the liquid not only changes with time but also changes along the axial direction z from the value 7; at the entrance to the value T: at the exit, We will assume that the temperature does not change along the radius of the pipe. Conse- quently, we have two independent variables, z and 1. The state variable of interest for the heat exchanger is the temperature T of the heated liquid. Therefore, we need the energy balance for the characterization of the temperature. To perform this balance, consider the element of length Az defined in Figure 4.9 by the dashed lines. For this system and over a period of time Ar, we have Energy balance: PA Bz (Ties (Td = peyvA(T)|2 Mt = popvA(T) esac At ‘accumulation of] [flow in of flow out of } enthalpy during | | enthalpy during || enthalpy during the time period | | the time period || the time period At Lae Loar J +Q At (xD Az) 4.7) catalysed | tom te seem tthe gud trough ie | sat arn the time period where Q = amount of heat transferred from the steam to the liquid per unit of time and unit of heat transfer area A = cross-sectional area of the inner tube v = average (assumed constant) velocity of the liquid D = external diameter of the inner tube Dividing both sides of eq. (4.17) by Az Ar and letting Az ~0 and Ar 0, Steam Thi | Tlevar - =a Seam re 4.9 Tubular heat exchanger. 70 Modeling the Dynamic and Static Behavior of Chemical Processes Part! we take ar or 22 5 pepvA Fo nD 4.18) Per Ti+ err = RDO (4.18) Im eq, (4.18) we can substitute Q by its equal, Q-U(Tx-T) and take ped ad + peyrd Uw xD UTu-T) (419) a a ‘This is the equation of state that models the behavior of liquid’s tempera- ture (state variable) along the length of the exchanger. Since ea. (4.19) isa partial differential equation we say that the exchanger has been modeled as a distributed parameter system. Note that U is the overall heat transfer coefficient between steam and the liquid in the tube, and 7. is the ‘temperature of saturated steam. Example 4.13: Mathematical Model of an Ideal Binary Distillation Column Consider a binary mixture of components A and B, to be separated into two product streams using conventional distillation. The mixture is fed in the column as a saturated liquid (ie. at its bubble point), onto the feed tray f (Figure 4.10), with a molar flow rate (mol/min) Frand a molar fraction of component A, cj. The overhead vapor stream is cooled and completely condensed, and then it flows into the reflux drum. The cooling of the overhead vapor is accomplished with cooling water. The liquid from the reflux drum is partly pumped back in the column (top tray, N) with a molar flow rate Fe (reflux stream) and is partly removed as the distillate product with a molar flow rate Fo. Let us call Mao the liquid holdup in the reflux drum and xp the molar fraction of component A in the liquid of the reflux drum. It is clear that xp is the composition for both the reflux and distillate streams. ‘At the base of the distillation column, a liquid product stream (the bottoms product) is removed with a flow rate Fz and a composition xs (molar fraction of A), A liquid stream with a molar flow rate Vis also drawn from the bottom of the column and after it has been heated with steam, it returns to the base of the column. The composition of the recirculating back to column stream is xp. Let Mz be the liquid holdup at the base of the column. The column contains N trays numbered from the bottom of the column to the top. Let M, be the liquid holdup on the /th tray. The vapor holdup on each tray will be assumed to be negligible. In Figure 4.11a we see the material flows in and out of the feed tray. ‘Similarly, Figure 4.11b and c show the material flows for the top (Nth) and bottom (first) trays, while Figure 4.1ld refers to any other tray. To simplify the system, we will make the following assumptions: Cchap.4 Development of a Mathematical Model n Treas J Figure 4.10 Binary distil jon column, 1. Vapor holdup on each tray will be neglected. 2. The molar heats of vaporzation ofboth components A and B are approximately equal Tht means that | mol of condensing Vapor A ee ey heat to vaporize 1 mol of liquid. reat losses from the column to the surt ire ‘ote wenlgios, roundings are assumed 4 The relative volatility af the two components remain 5 eer the column, pe * constant ich tray is assumed to be 100% efficient (ie., the ‘ efficient (i, the vapor levis each tray is in equilibrium with the liquid on the my). " ‘The first three assumptions imply that Ve Vin Vance ty and there no nee for energy bane around exch a. last two assumptions imply that a simple vapor-lguid eqilib- sum relntip canbe wed to tate the molar fection of An the ~aporeaving te th way (3) with the mola ration OFA vo te hal leaving the same tray (x;): orm in ine aug "Teta (4.20) where ais the relative volatility of the two components A and B. 72 Modeling the Dynamic and Static Behavior of Chemical Processes Parti ‘The final assumptions that we will make are the following: 6. Neglect the dynamics of the condenser and the reboiler. It is clear that these two units (heat exchangers) constitute processing systems ‘on their own right and as such they have a dynamic behavior (see Example 4.12). Therefore, accurate modeling should include the state equations, which describe the dynamic behavior of condenser and reboiler. 7. Neglect the momentum balance for each tray and assume that the molar flow rate of the liquid leaving each tray is related to the liquid holdup of the tray through the Francis weir formula: .N 42) Li=fM) i Ado he Let us now develop the state equations that will describe the dynamic behavior of a distillation column. The fundamental quantities are total ‘mass and mass of component A. But the question is: What is the system. ‘around which we will make the balances? From a practical point of view, Ym Feed way Ne tay o ith way @ @ Figure 4.11 Modeling details of the binary distillation column: (a) feed section; (b) top section and overhead accumulator, (c) bottom section land reboiler; (4) general ith tray chap. 4 Development of a Mathematical Model the boundary of the system of interest is outlined by dashed lines in Figure 4.10. Such a boundary clearly identifies the inputs and outputs of practical significance for the overall system. It is also evident that unless we can describe how the concentrations and liquid holdups on each tray cchange with time, we cannot find how the variables of practical signifi- cance, such as xp and xs, change with time. Therefore, we are forced to Gonsider the balances around each tray. Thus we have (se also Figure Feed tray (i a(My Total mass: MM = Fe Loy + Vou Lym Vom Ft Lys Ly (428) (Myx; Component a 1M Free Larnga+ Vestes— Lay Vor (422) Top tray (i= NY: Total mass: M6 54 Vy Le Vw Fr Ly (4238) > d(Myxy) Component A: per at = Faso + Vyaayyas= Low = Vag (4.230) Bottom tray (i = 1: M) Total mass: M9 15-2, 4V— Vim Labi (24a) Component a: WME) paxs+ Woy Lii~ Pir (A246) ith way (= 2,..., N= Land # #f win Lit Via- Vis Lib, (4.258) atx Component A: SM) Laasia + Hiayna~ Lani Vo, (4.256) Reflux drum: Total mass: 400) y, - Fy — Fy (4.260) aU Mrox0 Component A: SOfAD%) «Yay (+ Fo)xe (4.260) Column base: aM) Total mass: 4 1-V-Fe a) WO) 1 v—F, (4278) Component A: sites) Lixi~ Wye Fate (4.270) 73 74 Modeling the Dynamic and Static Behawor of Chemical Processes Parti All the equations above are state equations and describe the dynamic behavior of the distillation column. The state variables of the model are: Liquid holdups: M, May... Mp--. Mui Mro and My Liquid concentrations: X1, X24...» Xo» Xxi Xo and Xo jing of the column, in addition to the state mnships: To complete the modi equations, we need the following rel 1. Equilibrium relationships: ax, oj .N,B (420) wey (a= I)x, 2, Hydraulic relationships (Francis weir formula) L=fM) fA WeohowN O20 ‘When all the modeling equations above are solved, we find how the flow rates and concentrations of the two product streams (distillate, bot- tom) change with time, in the presence of changes in the various input variables. ‘The modeling steps outlined above indicate that the overall procedure may be tedious and full of simplifying assumptions. At times the resulting model is overwhelming in size and the solution of the corresponding equations may be cumbersome. For the binary distillation column we have to solve a system of 2N + 4 nonlinear differential equations (state equations) and ‘2N +1 algebraic equations (equilibrium, and hydraulic relationships) 4.6 Modeling Difficulties ‘The modeling examples discussed in previous sections of this chapter should have alerted the reader to a series of difficulties that we may encounter in our efforts to develop a meaningful and realistic mathe- matical description of a chemical process. Example 4.14: Difficulties in the Modeling of a CSTR Considering the mathematical modeling of a CSTR (Example 4.10), the following difficulties arise: 1. Determine with the desired accuracy the values of various parame- ters such as the preexponential kinetic constant ke, the activation energy E, and the overall heat transfer coefficient U. 2. although the specific heat capacities, cp and cp, have been consid ered constant, they are in general functions of the temperature T Cchap. 4 Development of a Mathematical Model 75 ‘and the concentration cy, How do we decide that this dependence is, weak (so that we can use constant values as in the example) or strong (in which case the modeling becomes very complicated)? ‘The same questions arise for the densities p and p, and the heat of | reaction (~AH,), 3. During the operation of the CSTR, scaling, fouling, and so on, will alter the value of the overall heat transfer coefficient, How can we account for this effect in the mathematical model? 4, We have considered first-order kinetics to describe the reaction rate. Is this correct? We can classify the difficulties encountered during the mathematical modeling of a process in three categories: 1. Those arising from poorly understood chemical or physical phe- nomena 2. Those caused from inaccurate values of various parameters 3. Those caused from the size and the complexity of the resulting model Poorly understood processes To understand completely the physical and chemical phenomena occurring in a chemical process is virtually impossible, Even an accept- able degree of knowledge is at times very difficult. Typical examples include: Multicomponent reaction systems with poorly known interactions among the various components and imprecisely known kinetics Vapor-liquid or liquid-liquid thermodynamic equilibria for mul- ticomponent systems Heat and mass transfer interactions in distillation columns with nonideal multicomponent mixtures, azeotropic mixtures, and so on. Example 4.15 Consider the fluidized catalytic cracking process shown in Figure 4.12. An oil feed composed of heavy hydrocarbon molecules is mixed with catalyst and enters a fluidized bed reactor. The long molecules react on the surface of the catalyst and are cracked into lighter product molecules (such as gasoline) which leave the reactor from the top. While cracking is taking place, carbon and other heavy uncracked organic materials are deposited on the surface of the catalyst, leading to its deactivation. The catalyst is then taken into a regenerator, where the material deposited on its surface is burned with air. Then, the regenerated catalyst returns to the reactor after it has been mixed with fresh feed, 76 Mexeing th Dyaric and Stave Behavior of Chea! Processes Part com 0 > ots othe ato { ms Deactivated cast a Reactor ae Rene catalyst Hiya feed Figure 4.12 Fluid catalytic cracking (FCC) system. ‘To model the two units, the following information must be available: 1. The reaction rate of the cracking process 2. The rate with which carbon and heavy material are deposited on the catalyst (this will determine the rate of catalyst deactivation) 3. The dependence of the two rates above on the temperature of the reactor and the quality of the feed (light or heavy) 4. The rate with which carbonaceous material deposited on the cata lyst is burned off in the regenerator, and its dependence on temper- ature i 10 acquire, but at Al of the foregoing information is not only difficult to acquire, times it leads to contradicting contentions. For example, in Figure 4.13 we see two models that describe the effect of the heavy oil feed rate on the reactor temperature. We notice that the qualitative behavior predicted by the two models is quite different. Finally, the two units (reactor, regenerator) are fluidized beds and it is ‘well known how poorly understood the fluid mechanical characteristics of such units are. Reactor Lee-Kugelman model Fede Figure 4.13 Two different models to describe the effect of heavy ol feed rate on reactor temperature for the FCC unit Cchap. 4 Development of a Mathematical Modal 7 Imprecisely known parameters The availability of accurate values for the parameters of a model is indispensable for any quantitative analysis of the behavior of a process. Unfortunately, this is not always possible. Typical examples include the preexponential constant of a kinetic rate expression. It should also be pointed out that the values of the parameters do not remain constant over long periods of time. Therefore, for effective modeling we need not only accurate values but also some quantitative description of how the parametric values change with time. Typical examples of changing parameters are the activity of a catalyst and the overall heat transfer coefficient of heat transfer systems (heat exchangers, jacketed reactors, etc.). The dead time is also a critical parameter whose value is usually imprecisely known and varying. As we will see in a later section, poor knowledge of the dead time can lead to serious stability problems for the process, When no reliable values for the parameters are available, we resort to experiments on the real process in an effort to estimate some “good” values for them. The experimental procedures will be discussed further in Chapter 31. Size and complexity of a model In an effort to develop as accurate and precise a mathematical model as possible, its size and complexity increase significantly. Example 4.16 Consider a distillation column with 20 trays, a reboiler, and a con- denser. The feed is a two-component mixture. Then, as we have seen in Example 4.13, the mathematical model is composed of 2N +4=2(20) +4 = 44 differential equations and 2N +1 =2(20) +1 = 41 algebraic equations The size of the model for such a simple system is already prohibitive. Since the common distillation systems include feeds with more than two components and possess larger numbers of trays, itis clear that such an. extensive modeling would lead to cumbersome and hard-to-use models. ‘Therefore, care must be exercised that the size and complexity of a model do not exceed certain manageable levels, beyond which the model loses its value and becomes less attractive. 78 Modeling the Dynamic and Static Behavior of Chemical Processes Pert THINGS TO THINK ABOUT What is a mathematical model of a physical process, and what do we mean when we talk about mathematical modeling? 2. In Figure 4.13 we see two different curves that relate the temperature and the feed rate of the reactor for the fluid catalytic cracking unit discussed in Example 4.15. Is the term “model” appropriate for each of these curves? 13, Let us recall that the steam tables give the temperature at which water liquid and water vapor are at equilibrium for a given pressure. They also give the specific values for enthalpy, entropy, and volume of both liquid and vapor phases. Do these tables of values constitute a mathematical model? 4, Consider the graphs shown in Figure Q4.1. These graphs were produced by ‘measuring the concentration of B in the reaction A ~ B, over time, and at ‘various temperatures. Do these graphs represent a mathematical model? Concentration of B r n Temperate Ty Time Figure Q4.1 ‘5, Why do you need to develop the mathematical model of a process you want to control? 6. What are the state variables, and what are the state equations? What are they used for? 7. How many state variables do you need to describe a system that is com- posed of M phases and NV components? 8. We know that when two phases are at thermodynamic equilibrium, the chemical potential j., of every component (i) in phase I is equal to the chemical potential 1 of the same component in phase I: i ee ee Express the equilibrium relationship above in terms of the mole concentra- tions of the NV components in the two phases. The answer to this question will demonstrate to you that we do not need the concentrations of the N components in both phases in order to describe the system. 9, Write a relationship that will give you the molar or the specific enthalpy of CChap.4 Development of a Mathematical Model 79 10. uu 2. 13. 4, 15. 16. 7. 19. 20. a. a multicomponent liquid at temperature 7" and pressure p, with known ‘composition for the N components. ® ” ° Repeat question 9, but with a gas instead of a liquid. |. Consider the flash drum of Examples 4.7 and 4.8, Develop an expression for the density of the vapor phase, using the van der Waals equation of state, State also an expression for the density of the liquid phase. ‘When is a system at steady state? What is the main reason for the presence of dead time in a process? Do you know of any systems that do not possess dead time? How would you find the dead time of a system? 4 Figure Qt.2 we see the behavior ofthe concentration a the outlet of two yrocesses after concentration at the inlets and at time ¢ = 0 was increased by 10%. Which process possesses dead time? Concentration 10 Time Figure Q42 What are the assumptions leading 10 equimolar vapor flow rates (ic., VieVi= Vy = V) fora binary distillation column? Why have we neglected the eneray balances forthe binary ideal sia column of Example 4.13? ann What are the assumptions kading to the enum rations and how is it derived? “a lationship (420) Could you nave dead time between the overhead vapor andthe ds Coul ou nave da por and the distillate Consider again Example 49, Show that the dead time can be com from the following equation . * pared [rt voane oe pe where F(0) is the volumetric flow rate of the li Ww rate of the liquid through the pipe as a function of time. The above equation is more general than that of Example 4.9, where the volumetric flow rate was assumed to be constant.

Вам также может понравиться