You are on page 1of 18

Preprint for Submission for Computer Applications in Engineering Education

Title: Practical Case Studies for Undergraduate Process Dynamics and Control Using the Process Control Modules (PCM)

Authors: Francis J. Doyle III, Edward P. Gatzke, Robert S. Parker

Author Affiliations: Department of Chemical Engineering University of Delaware Newark, DE 19716

All correspondence should be addressed to: Francis J. Doyle, email: doyle@che.udel.edu phone: (302) 831-0760, fax: (302) 831-0457

Practical Case Studies for Undergraduate Process Dynamics and Control Using the Process Control Modules (PCM)

Francis J. Doyle III, Edward P. Gatzke, Robert S. Parker

Department of Chemical Engineering University of Delaware Newark, DE 19716

Abstract: An effective environment for the incorporation of realistic case studies is introduced for a process dynamics and control course. The Process Control Modules (PCM) can be used in a

computer laboratory under MATLAB 5.1 and Simulink 2.1 (Mathworks). Several industrial case studies are reviewed which exemplify the use of such a tool for control education.

Keywords: Process Control Education, MATLAB/ SIMULINK, Chemical Engineering, Process Simulation

INTRODUCTION

The need for decreased variance in product quality coupled with tighter environmental regulation in the chemical process industries motivates improved process control of both existing processes and systems under development. Industrial demand for better process control implies that a detailed understanding of process control issues is expected of undergraduate students entering the workforce. Typical undergraduate process control classes are very mathematical, with lectures focusing mostly on the theory of modeling, design, and analysis. Many students find control theory concepts difficult to comprehend. A process control laboratory which can demonstrate and motivate the ideas presented in lecture is one effective means for transferring the knowledge to the students. There are many different objectives to be considered when developing a process control laboratory. Theoretical concepts should be presented in a realistic and understandable manner. Students should find the covered topics to be well motivated and clearly applicable in the real-world. A control laboratory with multiple unit operations controlled by a DCS would certainly demonstrate industrial principles and methods to the students, but this type of laboratory would be prohibitive to many universities because of economics, safety, and operation time constraints. Distance learning may also factor into the course laboratory objectives. A good solution for all of these concerns is the development of computer simulations which realistically demonstrate theoretical issues to students. There are other similar software packages which are currently available. Cooper [1], Koppel [2], Marlin [3], and Bequette [4, 5] have developed software packages for process control education. With the exception of [3] and [4, 5], these packages do not handle realistic large scale problems in an open programming environment (such as MATLAB).

PROCESS CONTROL MODULES

The Process Control Modules (PCM) package is comprised of many units, each covering a different topic in process dynamics and control. Most of the modules focus on the modeling or control of a typical chemical engineering unit operation. Currently, four different process models are available with PCM. These include a heated tube furnace, a binary distillation column, a continuous fermentation biological reactor, and a continuous Kraft pulp digester. In each of these cases, a detailed fundamental process description is employed These realistic case studies serve as a valuable component in the control curriculum.

PCM is a flexible package. Each module is largely independent of the others, so an instructor can elect to use a subset of the modules for assignments. New unit operation models can also be quickly added. For example, a company may want to use PCM to hold a process control course for employees using a model of their existing operation instead of the models provided with PCM. Advanced controller designs can be incorporated into PCM for use in graduate level control classes. PCM is also flexible in that it will run on any platform supported by MATLAB (Win 95, Win NT, Mac, UNIX). A manual is included with the PCM software. This manual provides information, detailed software directions, and student simulation exercises related to the module topic. The manual explains how to run the PCM software on a very detailed level, assuming that the students have no prior experience with MATLAB and SIMULINK. The manual contains basic information about different elementary process control topics, but it is not intended as a replacement for a standard textbook. Each laboratory module is subdivided into four sections: objective, introduction, exercises, and summary. In addition, each module includes a section entitled "Things to Think About" which is essentially a pre-lab, or set of exercises, for the students to complete before running the simulations. This section helps to orient students to the units topic and introduce concepts related to the current module. Short descriptions of the topics currently covered by the Process Control Modules are included in the following section. It is worth noting, that all of the units (except the Introduction and the First and Second Order Systems) deal with a specific case study. In the offerings of this laboratory by the first author at both Purdue University and the University of Delaware, several unit operations have been rotated into the computer laboratory on a annual basis. PCM Module Descriptions

Introduction. This laboratory module introduces the students to MATLAB and SIMULINK. Elementary command line exercises and block diagram simulations are performed. The students also learn how to start modules using the mainmenu graphical user interface (Figure 1). This module is independent of a particular unit operation.

Steady-State Gains. This is the first module to focus on a particular unit operation. Students change values for system inputs in a SIMULINK window. The system output levels are displayed in the simulation process monitor window (Figure 2). Students can use a click and point interface to ascertain process variable values from the strip chart windows. After collecting step change data, the students can then calculate steady-state gains for the unit operation. The operating point is then changed and gains are recalculated. This demonstrates to students the nonlinear effects typical of an actual process. The students also calculate and evaluate steady-state control moves, given a known disturbance and desired output change. First and Second Order Systems. In this module, students explore the dynamic characteristics of first and second order transfer functions. Students are presented a SIMULINK simulation window and a monitor which displays the system input and output values. From the time domain response to an input step change, the students calculate relevant system parameters. For first-order systems, this includes the gain and time constant. For second order systems, the students determine values for the gain, natural period, dampening coefficient, overshoot, and decay ratio. After exposure to the various effects of parameter variation, the students are presented with several unknown problems and are asked to identify the system parameters. While the processes employed in these units are low order (first and second order, e.g. surge tanks), the students are given a chance to familiarize themselves with the graphical nature of the SIMULINK interface before being introduced to the more complex case studies.

Transient Response. This unit demonstrates the development of transfer function models from step test data. The students collect data from step changes applied to one of the case studies. The data is displayed using PCMplot, a PCM-specific graphical tool for process modeling, identification and validation (Figure 3). This data is analyzed to develop first-order models with a time-delay. Due to the high order nature of the underlying nonlinear process dynamics and the random noise signal applied to the process measurements, the students are exposed to the kind of realistic data they would encounter in practice at a chemical plant, refinery, mill, etc.. Hence, they will find that a first-order (plus delay) model is very simple from a computational perspective, but it is at best an approximation to a real system. In the course of the model identification, students develop a transfer function simulation in SIMULINK. Given the original process data, and data generated by their identified model, students can validate the model fidelity using PCMplot.

Open Loop Frequency Response. Frequency response techniques can give students powerful tools for use in industry. In this module, students learn how to create a frequency response model. Classic textbook treatment of this problem uses sinusoidal forcing of the process to generate individual frequency points, however this is hardly ever done in practice. The more practical approach involves a transformation of time-domain (discrete) data into frequency domain data. This is accomplished using pulse-testing along with a fast-fourier transformation. The pcmfft command performs the fourier transform, and the plot_bode command creates a Bode plot (Figure 4) with chemical engineering units. Initially, the students create bode plots from first and second-order transfer functions. The same exercise is repeated using the unit operation model. As with the previous identification unit, the students encounter all the practical limitations of finite magnitude inputs, finite duration signals, noise-corrupted data, and complex dynamic process behavior. The Bode plots which result from the case study do not display the nice properties of low order systems (as seen in the slope of the high frequency asymptote of the amplitude ratio). Therefore the students must exercise engineering judgment in developing an approximate model that is consistent with their data and is suitable for subsequent controller design.

Steady State Analysis of Feedback Control . This laboratory module introduces the concept of feedback control. The students use the models generated in the transient response module to develop PID tuning constants using standard tuning rules. Using setpoint changes and disturbance loading, the students examine the effect of integral action on steady-state system control.

PID Controller Tuning. This unit allows the students to explore the dynamic response of a process case study using PID feedback controllers. The students introduce step changes and disturbances to the system and monitor the output variable response. Using the models developed earlier (transfer function and frequency domain), the students apply three distinct tuning algorithms and compare the results. These algorithms include Cohen-Coon, Ziegler-Nichols, and relay tuning. The latter tuning approach is important as it represents a modern approach to control tuning which is available in many commercial control packages, but is only treated in the most current undergraduate textbooks. Its efficacy for a complex system, for which the other two tuning methods fail, is highlighted in this unit.

Feed-Forward control. Using the process models developed in the transient response module, students develop, implement, and test feed-forward control designs. Again, the principle of approximation is reinforced, as the students learn that simple lead-lag feed-forward controllers do not exactly cancel the responses of complex systems. However, the students find that even simple

approximations can be highly effective in the anticipatory response provided by feed-forward control.

IMC and Multivariable Control. In this unit, the system process model is used to develop an Internal Model Control (IMC) controller. This module highlights the role of advanced model-based design methods for controller synthesis. Once again relatively few textbooks address this approach at the undergraduate level, but the concepts are straightforward, and the application in the laboratory can significantly reinforce the learning. The tuning procedure is reduced to two intuitive time constants (one per loop), and the students find that low order model approximations can serve as valuable components in a direct synthesis controller design method. The second half of this lab introduces multivariable control design using a decoupling approach.

Model Predictive Control. This module demonstrates Model Predictive Control (MPC), also known as Dynamic Matrix Control (DMC). The transfer function models obtained in earlier modules are discretized to develop a discrete step-response model. This controller design methodology is considered by many to be the current state of the art in advanced control design, and it gives the students a brief introduction to the important issues. It also raises the important issue of digital computing, and the role of optimization in solving control problems. The tuning procedure becomes more complicated, as it would for any realistic case study. The students are challenged to find suitable values for the horizons used in MPC (move and prediction), as well as the weighting matrices (input and output).

PCM Unit Operations

Typical Unit. Any process unit operation with a continuous finite state space realization can be modeled in the SIMULINK environment. The PCM laboratory modules currently make use of four fundamental models, but new modules can easily be developed for use with different unit operations. Ideally, a new unit operation would have at least four inputs and four outputs. Two of the inputs should be manipulated variables and the other two should be disturbance variables. For feed-forward control purposes, one of the disturbance inputs should be a variable that could be realistically transmitted by a field sensor. The SIMULINK block for any new unit operation should use absolute process variables for inputs and outputs. A process model should take less than ten actual minutes to simulate an input step change response, otherwise its value in the laboratory classroom will be limited. The student interface to a typical unit is shown in Figure 5, where it is evident that the student has convenient access during run-time to parameter values in the process (inputs, tuning parameters, etc.) through a convenient graphical environment.

There are different methods available for developers to create unit operations models in the SIMULINK environment. One may model a system in the form:

where x is a vector of continuous states, u is the vector of inputs, and y is the vector of outputs. Templates are available in SIMULINK for implementing such a set of equations using FORTRAN, C, or MATLAB compiler languages. Note that the use of the FORTRAN or C languages usually improves the simulation speed up to fivefold in our experience. The FORTRAN or C development method requires use of an external language compiler working with MATLAB, and the compiled models will not be useable in the student version of MATLAB. Another modeling option is the Differential Equation Editor (DEE) delivered with the SIMULINK 2.1 environment. This allows the user to easily create a system model from the functions f , g, system parameters, and initial states.

Furnace. This model was originally developed in 1982-1983 under the direction of Professor Lowell Koppel for use in the IBM Advanced Control System (ACS) instructional modules at Purdue University. The system models the heating of a high-molecular weight hydrocarbon feed to a cracking unit of a refinery. An air-fuel mixture undergoes combustion in the furnace to heat the hydrocarbon feed. The furnace model has 26 continuous states, with a detailed description given in Appendix A. The furnace has two manipulated process inputs: air flow rate and fuel gas flow rate. There are five disturbance inputs: hydrocarbon inlet temperature, hydrocarbon flow rate, air temperature, fuel gas temperature, and fuel gas purity. The measured process variables are hydrocarbon outlet temperature, furnace temperature, exhaust gas flow rate, and oxygen exit temperature.

This unit is particularly effective in demonstrating the effect of transport delay through a distributed parameter system, and the disparate time scales associated with mass and energy effects in a reactor.

Binary Distillation Column. This unit operation models the separation of a mixture of ethanol and methanol. The column has 27 trays with a reboiler on the bottom tray and a total condenser on the overhead stream. This model was originally developed at the University of Maryland [6]. Manipulated process inputs are reflux ratio and vapor flow rate. The feed methanol mole fraction and feed flow rate act as disturbance inputs. The measured process outputs are: overhead methanol mole fraction, overhead product flow rate, bottom methanol flow rate, and bottom product flow rate. This model has 60 continuous states.

This unit is most effective for demonstrating the interactive nature of multivariable processes, as well as the nonlinear nature of high purity columns in the limit of high reflux or vapor boilup.

Biological Reactor. This unit operation is a seven state model of a continuous flow fermentation biological reactor. A biological organism (SACCHAROMYCES CEREVISIEA) is contained in a CSTR with a glucose solution feed stream. The model is taken from [7, 8]. There are seven states in the model. The manipulated process inputs are dilution rate and glucose feed concentration. Disturbance inputs are specific glucose uptake rate and growth rate. The outputs are cell biomass concentration, exit glucose concentration, and ethanol concentration.

This unit represents one of the more modern unit operations that have appeared in the chemical engineering curriculum. It can be easily modified to accommodate other organisms and metabolic pathways.

Pulp Digester. This process is the key unit operation at the front end of a pulp mill, in which the wood chips are broken down using a caustic liquor. The control objective involves the regulation of the final pulp properties, despite upsets in the biological feedstock properties. The computer model is derived from [9, 10] and employs over 1000 dynamic states in representing the distributed parameter behavior through a lumped (series of CSTRs) approximation. The inputs include upper extract flowrate, cooking temperature, MCC temperature, MCC trim flowrate, chip flowrate, and chip moisture. The outputs are kappa number, upper and lower effective alkali, and lower Na2S concentration.

This unit exposes students to another sector of the industry that is often under-represented in the chemical engineering curriculum, but is a large employer of chemical engineers the pulp and paper industry. Furthermore it highlights the challenges of working with long time delay systems (~ 6-8 hours).

FUTURE WORK

There are a number of ongoing efforts to develop and improve the PCM package. In the next year, there will be two notable enhancements to the graphical interface. The MATLAB GUI design tool (GUIDE) will be used to reformat the process monitor windows, and provide easier user access to simulation parameters. PCM development is also exploring the use of multimedia. This will make PCM a more powerful learning experience by providing audio and visual instruction to supplement the current workbook and simulation. The new multimedia content of PCM will offer information and explanations about different process control and PCM topics. The content will also allow the development of a plant-like operating environment. For example, this may be accomplished by providing a video of an operator adjusting a valve in a plant or creating warning bells when process constraints are violated. The interactive experience will help to capture the attention of the student, increasing the retention and understanding of the material.

CONCLUSIONS

The Process Control Module system provides students with an opportunity for a hands-on, realistic, computer laboratory experience. Use of fundamental models exposes students to conditions that would be observed in actual process systems, such as nonlinear system gains, high order dynamics and noise-corrupted measurements. MATLAB and SIMULINK allows PCM to support any MATLAB supported platform (Win95, WinNT, Mac, UNIX). The SIMULINK environment allows for easy development, modification, and extension of the current package.

ACKNOWLEDGMENTS The authors acknowledge the support of the following students who have helped develop the PCM software: Atsushi Aoyama, Arda Bafra, Lalitha Balasubramhanya, Derek Brown, Clark Case, Jorge Castrove, Matthew Clamme, Joseph Cooper, Maruti Dey, Edward Gatzke, Jason Gause, Douglas Heemstra, Steven Honkomp, Jeffrey Kao, Thomas Kendi, Daniel King, Harpreet Kwatra, Alan

Mahoney, Bryon Maner, Kerry Need, Robert Parker, Anh Phung, Kairali Podual, Andre Shaw, Alexander Stack, Phillip Wisnewski, and Stanley Wrobel.

ADDITIONAL INFORMATION ABOUT PCM

Interested educators are referred to the Web-page for the Process Control Modules at: http://www.che.udel.edu/pcm

Appendix A Furnace Details

The following 2 combustion reactions are modeled in the furnace:

Reaction 1:

Reaction 2:

The specific mass and energy balance equations for this unit operation are detailed below. The equations are given in a MATLAB m-file s-function format. The percent symbol denotes code comments.

SYSTEM PARAMETERS:

% General Furnace Parameters effurn = 0.8712; % Energy Efficiency cpfurn = 1.05e+05; % Heat Capacity (J/K) vol = 250; % Reaction Volume (m3)

% Fluid Tubing Specifications n = 20; % Number of Partition L = 40; % Length (m) dL = L / n; % Length of Partion (m)

d = 0.08; % Diameter (m) h = 400; % Heat Transfer Coefficient (J/m2-min-K) csa = pi * (d / 2)^2; % Cross-Sectional Area (m2)

% Heated Fluid (Eicosane n-C20H42 ) rhofl = 786.6; % Density (kg/m3) cpfl = 463.31; % Heat Capacity (J/kg-K) tauh = rhofl * cpfl * d / (4 * h); k1 = 400; % Rxn 1 Rate Constant (m^4.5/mol^1.5-s) k2 = 200; % Rxn 2 Rate Constant (m^1.5/mol^0.5-s) hrxn1 = 6.0443e+05; % Rxn 1 Heat of Reaction (J/mol) (T = 1400K) hrxn2 = 2.8498e+05; % Rxn 2 Heat of Reaction (J/mol) (T = 1400K)

% Initial Heat Capacities (J/mol-K) (T = 310K) cpch4i = 35.826; % Methane cph2oi = 33.634; % Water cpo2i = 29.595; % Oxygen cpn2i = 29.144; % Nitrogen cpcoi = 29.235; % Carbon Monoxide cpco2i = 38.053; % Carbon Dioxide % Final Heat Capacities (J/mol-K) (T = 1400K) cpch4f = 84.586; % Methane cph2of = 45.778; % Water cpo2f = 36.048; % Oxygen cpn2f = 34.189; % Nitrogen cpcof = 34.538; % Carbon Monoxide cpco2f = 57.042; % Carbon Dioxide

SYSTEM INPUTS:

% u(1) = Hydrocarbon Fluid Flow Rate (m3/min), ss=0.035 % u(2) = Hydrocarbon Inlet Temperature (K), ss=310 % u(3) = Air Flow Rate (m3/min), ss=17.9 % u(4) = Air Temperature (K), ss=310 % u(5) = Fuel Gas Flow Rate (m3/min), ss=1.21 % u(6) = Fuel Gas Temperature (K), ss=310 % u(7) = Fuel Gas Purity (moles CH4/mole), ss=1

ALGEBRAIC AND STATE EQUATIONS:

rhoig = P / (R * u(4)); % Ideal Gas Density (mol/m3) feed = u(3) + u(5); % Combustion Reactants Feed Rate (m3/min) velfl = u(1) / csa; % Velocity of Fluid (m/min)

% Initial Concentrations (mol/m3) cch4i = u(7) * u(5) * rhoig / feed; % Methane ch2oi = (1 - u(7)) * u(5) * rhoig / feed; % Water co2i = 0.21 * u(3) * rhoig / feed; % Oxygen cn2i = 0.79 * u(3) * rhoig / feed; % Nitrogen ccoi = 0; % Carbon Monoxide cco2i = 0; % Carbon Dioxide

% Initial Species Flow Rates (mol/s) nch4i = cch4i * feed; % Methane nh2oi = ch2oi * feed; % Water no2i = co2i * feed; % Oxygen nn2i = cn2i * feed; % Nitrogen ncoi = 0; % Carbon Monoxide nco2i = 0; % Carbon Dioxide

% Final Species Flow Rates (mol/s) nch4f = nch4i - k1 * x(1) * x(3) ^ 1.5 * vol; % Methane nh2of = nh2oi + 2 * k1 * x(1) * x(3) ^ 1.5 * vol; % Water

no2f = no2i - 1.5 * k1 * x(1) * x(3) ^ 1.5 * vol - 0.5 * k2 * x(4) * x(3) ^ 0.5 * vol; % Oxygen nn2f = nn2i; % Nitrogen ncof = k1 * x(1) * x(3) ^ 1.5 * vol - k2 * x(4) * x(3) ^ 0.5 * vol; % Carbon Monoxide nco2f = k2 * x(4) * x(3) ^ 0.5 * vol; % Carbon Dioxide ntotf = nch4f + nh2of + no2f + nn2f + ncof + nco2f; % Total

% Combustion Products Exit Flow (m3/min), from mass balance on carbon exit = (2 * no2i + nh2oi) / (x(2) + 2 * x(3) + x(4) + 2 * x(5));

% Mass Balance of Combustion Species

% Species Concentrations: % Methane x(1), Water x(2), Oxygen x(3), Carbon Monoxide x(4), Carbon Dioxide x(5) dx(1) = (1 / vol) * (nch4i - exit * x(1)) - k1 * x(1) * x(3) ^ 1.5; dx(2) = (1 / vol) * (nh2oi - exit * x(2)) + 2 * k1 * x(1) * x(3) ^ 1.5; dx(3) = (1 / vol) * (no2i - exit * x(3)) - 1.5 * k1 * x(1) * x(3) ^ 1.5 - 0.5 * k2 * x(4) * x(3) ^ 0.5; dx(4) = (1 / vol) * (ncoi - exit * x(4)) + k1 * x(1) * x(3) ^ 1.5 - k2 * x(4) * x(3) ^ 0.5; dx(5) = (1 / vol) * (nco2i - exit * x(5)) + k2 * x(4) * x(3) ^ 0.5; % Energy Balance of Combustion Species and Fluid, dT/dt for initial partition dx(1+5) = (velfl / dL) * (u(2) - x(1+5)) + (1 / tauh) * (x(5+n+1) - x(1+5));

% dT/dt for 2nd to nth Partition of Tubing (x(2+5) to x(n+5)) (K/s) for i = (2+5):(n+5) dx(i) = (velfl / dL) * (x(i-1) - x(i)) + (1 / tauh) * (x(5+n+1) - x(i)); end % Heat Exchange (J/s) % Energy Lost to Species as Enthalpy, (Final - Initial) {-} qi = (nch4i * cpch4i + nh2oi * cph2oi + no2i * cpo2i + nn2i * cpn2i) * u(4); qf = (nch4f *cpch4f+nh2of *ph2of+no2f*cpo2f+nn2f*cpn2f+ncof*cpcof+nco2f*cpco2f)*x(5+n+1); % Energy Generated by Reactions {+} qrxn1 = effurn * (k1 * x(1) * x(3) ^ 1.5 * vol) * hrxn1; qrxn2 = effurn * (k2 * x(4) * x(3) ^ 0.5 * vol) * hrxn2;

% Energy Lost to Fluid {-} qfluid = h * d * dL * pi * (n * x(5+n+1) - sum(x(5+1:5+n)));

% Energy Balance of Furnace (x(5+n+1), furnace temperature) dx(5+n+1) =(qrxn1 + qrxn2 - (qf - qi) - qfluid) / cpfurn;

References

[1] D. J. Cooper. "Picles : a simulator for virtual world education and training in process dynamics and control," Computer Applications in Engineering Education, Vol. 4, No. 3, 1996, pp. 207-215.

[2] L. B. Koppel and G. R. Sullivan. "Use of IBMs advanced control system in undergraduate process control education," Chemical Engineering Education, Vol. 20, 1986, p 70.

[3] T. E. Marlin. "Software laboratory for undergraduate process control education," Computers & Chemical Engineering Proceedings of the 6th European Symposium on Computer Aided Process Engineering. Part B. May 26-29 1996, Vol. 20.

[4] B. W. Bequette, K. D. Schott, V.Prasad, V. Natarajan and R.R. Rao. "Case Study Projects in an Undergraduate Process Control Course," Chemical Engineering Education (in press, 1998).

[5] B. W. Bequette, "Case Study Projects in an Undergraduate Process Control Course." Proceedings of Control-97, Sydney, p 212-217 (1997).

[6] K. Weischedel and T. J. McAvoy. "Feasibility of decoupling in conventionally controlled distillation columns," Independent Engineering Chemical Fundamentals, Vol. 19, 1980, p 379.

[7] Lievense, Jefferson Clay. "An investigation of the aerobic, glucose-limited growth and dynamics of SACCHAROMYCES CEREVISIEA," PhD Thesis, Purdue University, 1984.

[8] Aoyama Atsushi. "Modeling and control of nonlinear processes using neural networks and fuzzy

logic," PhD Thesis, Purdue University, 1994.

[9] P. A. Wisnewski. "Inferential control using high order process models with application to a continuous pulp digester," PhD Thesis, Purdue University, 1997.

[10] F. Kayihan. "Continuous Digester Benchmark Dynamic Model," Matlab Toolbox Code, IETek, 1997.

Figure 1: Main menu for Process Control Modules (PCM).

Figure 2: Process monitor for furnace unit, showing capability for data archiving, window control, axis rescaling, and pointer function (with depicted cross-hair).

Figure 3: Interface for PCMPlot utility, showing original process data (yellow solid line) and fitted first-order-plus-time-delay response (blue dash-dot line).

Figure 4: Output of plot_bode command, showing the amplitude ratio (top) and phase lag (bottom) over the frequency range of interest.

Figure 5:

Process flowsheet window showing the student interface for real-time operation of the furnace

You might also like