ap-4-12.dvi Acta Polytechnica Vol. 52 No. 4/2012 An Application of the Individual Channel Analysis and Design Approach to Control of a Two-input Two-output Coupled-tanks System David J. Murray-Smith1 1 School of Engineering, Rankine Building, University of Glasow, Glasgow G12 8AT, Scotland, United Kingdom Correspondence to: David.Murray-Smith@glasgow.ac.uk Abstract Frequency-domain methods have provided an established approach to the analysis and design of single-loop feedback control systems in many application areas for many years. Individual Channel Analysis and Design (ICAD) is a more recent development that allows neo-classical frequency-domain analysis and design methods to be applied to multi-input multi-output control problems. This paper provides a case study illustrating the use of the ICAD methodology for an application involving liquid-level control for a system based on two coupled tanks. The complete nonlinear dynamic model of the plant is presented for a case involving two input flows of liquid and two output variables, which are the depths of liquid in the two tanks. Linear continuous proportional plus integral controllers are designed on the basis of linearised plant models to meet a given set of performance specifications for this two-input two-output multivariable control system and a computer simulation of the nonlinear model and the controllers is then used to demonstrate that the overall closed-loop performance meets the given requirements. The resulting system has been implemented in hardware and the paper includes experimental results which demonstrate good agreement with simulation predictions. The performance is satisfactory in terms of steady-state behaviour, transient responses, interaction between the controlled variables, disturbance rejection and robustness to changes within the plant. Further simulation results, some of which involve investigations that could not be carried out in a readily repeatable fashion by experimental testing, give support to the conclusion that this neo-classical ICAD framework can provide additional insight within the analysis and design processes for multi-input multi-output feedback control systems. Keywords: multivariable, feedback, control, coupled tanks, frequency domain, nonlinear. 1 Introduction Problems of liquid level control arise in many indus- tries. Common examples include control of levels in blending and reaction vessels within chemical pro- cesses. This paper describes the application of the Individual Channel Analysis and Design (ICAD) ap- proach to the development, implementation and test- ing of conventional diagonal controllers for a system involving liquid levels in a pair of coupled tanks [1]. The objective of the paper is to provide a detailed case-study to illustrate use of the ICAD methodol- ogy and to demonstrate some of the benefits of this neo-classical frequency-domain approach to problems involving multivariable control. The coupled-tanks equipment around which the control system is designed has two inputs, which are external liquid flow-rates into each of the tanks. The two outputs are the resulting levels of liquid in the tanks. There is only one outflow and this is from the second tank. Figure 1 is a schematic diagram of this system. Established frequency-domain methods, such as Bode/Nyquist analysis, are of central importance as Figure 1: Schematic diagram of the coupled-tanks equipment a basis for design tools for single-input single-output feedback control systems in many application areas. The success of the frequency-domain approach is due, in part, to the graphical nature of these techniques which provides transparency and flexibility in satisfy- ing design specifications in the presence of practical constraints. The extension of classical methods of analysis and design to multivariable systems involv- 121 Acta Polytechnica Vol. 52 No. 4/2012 ing more than one input and more than one output can introduce difficulties. It may still be possible, in cases where cross-coupling is not strong, to design the control system using approaches involving one loop at a time. However, in cases where dynamic interac- tions between loops are significant, more skill and ex- perience is necessary in order to produce a successful design, and a process of tuning using trial-and-error methods may be needed. The Individual Channel Analysis and Design (ICAD) methodology was developed in the early 1990s and allows frequency-domain methods to be applied to the problems of analysis and design of multi-input multi-output feedback control systems (see, e.g., [2–6]). This approach allows an m-input m-output feedback control problem to be split into m single-input single-output problems without loss of structural information. Each controlled output is paired with a specific reference input to form what is termed a “Channel”. The ICAD approach makes direct use of the customer performance specifica- tion for different channels to provide a framework within which classical single-input single-output con- trol engineering concepts can be extended to multi- input multi-output cases involving significant levels of cross-coupling. Traditional methods for appli- cations involving single-input single-output (SISO) systems can thus be applied to multi-input multi- output (MIMO) problems. This includes the use of open-loop system information in the form of Nyquist and Bode plots for system analysis and design, along with simple measures of robustness, such as gain and phase margins. It must be emphasized that the ICAD approach is not, itself, a design method but should be viewed as a framework through which useful insight may be gained about the dynamics of the plant and the char- acteristics of the complete controlled system. Perfor- mance can be assessed for any chosen form of linear controller (which may be designed using any suit- able approach), and limitations of a design can be investigated. Thus, compared with some other ap- proaches to multivariable control, it can be based on traditional analysis and design methods familiar to all control system designers. Each channel has its own customer-defined performance specifications and these may be expressed in a simple way in terms of SISO requirements. 2 The coupled-tanks system The two-input coupled-tanks laboratory system of Figure 1 consists of a container (of volume 6 liters), with a central partition which divides it into two sep- arate tanks. Coupling between these tanks is pro- vided by a number of holes of various diameters po- sitioned near the base of the partition. The strength of coupling may be adjusted through the insertion of plugs into one or more of these holes. The system is equipped with a drain tap which is under man- ual control and this allows the output flow rate from one of the tanks to be adjusted. Both tanks have inflows from electrically driven variable-speed pumps and are equipped with sensors that can detect the level of liquid and provide a proportional electrical output voltage. The hardware is based around a single-input com- mercial product intended for teaching applications (TecQuipment Ltd) [1]. This had a flow input only to Tank 1 and was modified at the University of Glasgow through the addition of the second pump to provide the inflow to Tank 2. The original resistive level sensors have been replaced using more accurate differential-pressure based depth sensors. The derivation of a detailed nonlinear model of the system may be found in [7] and in a more recent conference paper [8] which also includes a very brief account of the application of the ICAD approach to the design of a control system for this process. The model is based on the application of the prin- ciple of conservation of mass to the liquid within each tank. Bernoulli’s equations provide the basis for de- termining the flow from one tank to the other and from the second tank to the external environment. This leads to the following pair of equations, involv- ing variables shown in Figure 1: A1 dH1 dt = Qi1 − Cd1a1 √ 2g(H1 − H2) (1) A2 dH2 dt = Qi2 + Cd1a1 √ 2g(H1 − H2) − (2) Cd2a2 √ 2g(H2 − H3) These equations describe the dynamics of the cou- pled tanks system, in nonlinear form, for all cases for which the level in Tank 2 is below that in Tank 1. It is, of course, possible to derive a similar set of nonlinear equations to describe the system for cases involving a liquid level in Tank 2 which is greater than the level in Tank 1. Parameter values for the laboratory system are as follows: Cross-sectional area of tanks, A1 = A2 = 9.7 × 10−3 m2. Cross-sectional area of the orifice between Tank 1 and Tank 2, a1 = 3.956 × 10−5 m2. Cross-sectional area of the orifice representing the outlet drain of Tank 2, a2 = 3.85 × 10−5 m2. Height of the outlet drain above the base of Tank 2, H3 = 0.03 m. Gravitational constant, g = 9.81 ms−2. Maximum flow rates for inputs to Tank 1 and Tank 2: Qi1 max = Qi2 max = 5 × 10−5 m3 s−1. 122 Acta Polytechnica Vol. 52 No. 4/2012 The minimum flow rates for these inputs are zero since the pumps are not reversible and thus negative inputs are not possible. Maximum levels of liquid in Tank 1 and Tank 2: H1 max = H2 max = 0.3 m. The minimum level possible in each tank is 0.03 m which corresponds to the height of the outlet drain. Values of the discharge coefficients Cd1 and Cd2 have to be determined empirically and the values ob- tained depend on the system operating point. The value used for Cd1 in the design calculations was 0.63 and the value used for Cd2 was 0.58. In addition to the above parameters, electrical signals in the system are related to the variables of the model (as described by Equations (1) and (2)) through the following two parameters: Pump flow-rate calibration constant, Gp = 7.2 × 10−6 m3s−1V−1. Liquid depth sensor calibration constant, Gd = 33.33 Vm−1. It should be noted that no dynamic representa- tion is included for the two pumps and the associ- ated electrical drives as it was found, through hard- ware testing, that they show a very fast response to changes of electrical input compared with the level changes within the tanks themselves. Time constants that are associated with the pumps were therefore ne- glected for the purposes of the control system design. For the preliminary stages of the design it is also appropriate to consider a linearised model, within which the variables represent small variations of sys- tem variables about steady state values. h̃1(t) = H̄1 − H1(t) (3) h̃2(t) = H̄2 − H2(t) (4) qi1(t) = Q̄i1 − Qi1(t) (5) qi2(t) = Q̄i2 − Qi2(t) (6) q23(t) = Q̄23 − Q23(t) (7) In Equations (3)–(7) the variables that have a hori- zontal bar above them denote values at the chosen op- erating point, which is normally defined by a steady- state condition. If Equations (1) and (2) are re-arranged in the standard state-space form, we get a pair of nonlinear equations: dH1 dt = f1(H1, H2, Qi1) (8) dH2 dt = f2(H1, H2, H3, Qi2) (9) Then, since the level H3 may be assumed con- stant, linearisation produces the standard linear state space model:⎡ ⎢⎣ dh̃1 dt dh̃2 dt ⎤ ⎥⎦ = ⎡ ⎢⎣ ∂f1 ∂H1 ∂f1 ∂H2 ∂f2 ∂H1 ∂f2 ∂H2 ⎤ ⎥⎦ [ h̃1 h̃2 ] + (10) ⎡ ⎢⎣ ∂f1 ∂Q1 ∂f1 ∂Q2 ∂f2 ∂Q1 ∂f2 ∂Q2 ⎤ ⎥⎦ [ qi1 qi2 ] In Equation (10) all the partial derivatives must be evaluated at the operating point (H̄1, H̄2, Q̄i1, Q̄i2). The resulting linearised equation, after evaluation of the partial derivatives, has the form: [ ˙̃ h1 ˙̃ h2 ] = ⎡ ⎢⎣ −α1 A1 α1 A1 α1 A2 −(α1 + α2) A2 ⎤ ⎥⎦ [ h̃1 h̃2 ] + (11) ⎡ ⎢⎣ 1 A1 0 0 1 A2 ⎤ ⎥⎦ [ qi1 qi2 ] where α1 = Cd1a1 2 √ 2g H̄1 − H̄2 (12) and α2 = Cd2a2 2 √ 2g H̄2 − H3 (13) The individual block transfer functions that de- scribe the plant in Figure 1 may then be derived from Equation (11) and are as follows: g11(s) = (α1+α2) α1α2 [ 1 + s A2 (α1+α2) ] 1 + (A1α1+A1α2+A2α1) α1α2 s + A1A2 α1α2 s2 (14) g21(s) = 1 α1 1 + (A1α1+A1α2+A2α1) α1α2 s + A1A2 α1α2 s2 (15) g12(s) = 1 α2 1 + (A1α1+A1α2+A2α1) α1α2 s + A1A2 α1α2 s2 (16) g22(s) = 1 α2 (1 + sA1 α1 ) 1 + (A1α1+A1α2+A2α1) α1α2 s + A1A2 α1α2 s2 (17) 3 An outline of the ICAD approach A linear time-invariant MIMO plant may be mod- elled using a transfer function matrix G. If a control matrix K is positioned in the forward path in cas- cade with the plant transfer function matrix G and immediately before it, a feedback loop can be created around the combined system described by the prod- uct KG. The essential feature of the ICAD approach is that loops are considered individually, by opening one loop while all other loops remain closed. Details of the ICAD methodology and applica- tions that have been considered previously may be found in papers by Leithead and O’Reilly (e.g. [2–4] 123 Acta Polytechnica Vol. 52 No. 4/2012 and [5]), who were responsible for the initial develop- ment of the approach. A bibliography of published papers and reports relating to ICAD methods has been made available by Kocijan [6]. The ICAD methodology allows a controller to be assessed in a very direct fashion, for a given plant and given design specifications, in terms of performance and in terms of compromises and possible trade-offs. The design goals typically may involve: 1. Steady state response 2. Transient response 3. Disturbance rejection 4. Closed-loop stability 5. Robustness to changes in plant characteristics 6. Protection of actuators from high-frequency sig- nals that might lead to excessive wear In the ICAD approach the significance of the ‘struc- ture’ of the plant in translating the given MIMO sys- tem into the equivalent set of channels is given special emphasis [2]. Figure 2: Block diagram of a general two-input two- output closed-loop system of the type being consid- ered in this application. (Adapted from a diagram in [2]) As is clear from the plant model, the coupled- tanks application described in this paper involves a two-input two-output system with feedback involv- ing two channels. Figure 2 is a block diagram that illustrates the type of system being investigated. If we consider the forward signal transmission from the reference signal r1 to the associated output y1, it may be seen that the signal follows two pathways. One path involves a direct link through the block g11 and the other is through the blocks g21, g12 and a block involving k2 and its associated feedback loop through g22. This diagram may be simplified to give the struc- ture shown in Figure 3 for the Channel C1. From considerations of symmetry, the Channel C2 may be handled in the same way to produce the simplified block diagram of Figure 4. Figure 3: Block diagram for Channel 1. (Adapted from a diagram in [2]) These block diagrams can be used to show that, ignoring the disturbance signal, each channel can be described using a single-input single-output transfer function: C1 = k1g11(1 − γh2) (18) and C2 = k2g22(1 − γh1) (19) where γ = g12g21 g11g22 (20) h2 = k2g22 1 + k2g22 (21) and h1 = k1g11 1 + k1g11 (22) Figure 4: Block diagram for Channel 2. (Adapted from a diagram in [2]) In Equations (18)–(22) and in Figures 3 and 4, the effects of coupling are represented as additive distur- bance terms at the outputs of each channel and this does not involve any loss of information. However, it must be emphasised that ICAD is not a single-loop design method since loop interactions are preserved. It can be shown (e.g., [2, 4]) that, for robustness to parameter uncertainties of the closed-loop system stability, the Nyquist plots of (1−γh1) and (1−γh2) must not lie close to the origin of the polar plane at frequencies near to or below the open-loop gain cross- over frequency. Hence, if the corresponding plots of γh1(jω) or γh2(jω) come close to the point (1, 0) in 124 Acta Polytechnica Vol. 52 No. 4/2012 the polar plane the conventional SISO gain margins for the effective transfer functions of C1 and C2 do not provide robust measures of stability. In such a case it may not be appropriate to attempt to use the ICAD approach unless some form of pre-compensator is introduced to modify the plant characteristics in an appropriate way [5]. It can also be stated [4] that the quantities h1(jω) and h2(jω) have magnitude values that are generally close to one below the gain crossover frequency, and the quantity γ(jω), which is termed the multivariable structure function, provides a measure of the strength of any inter-loop coupling in the system and can in- dicate whether or not this is benign. For the case of a two-input two-output system there is only one multi- variable structure function. However, in general, for systems having a larger number of input-output pairs there will be more than one structure function. In all cases it can be stated that when a mul- tivariable structure function is small the interaction effects are small. In the two-input two-output case, if the multivariable structure function is small over the complete frequency range of interest, the two channels behave, more or less, as two independent loops. On the other hand, if the multivariable struc- ture function is shown to have a large magnitude, at a frequency within the range that is important for the application being considered, the loop interac- tions become significant [2]. If the multivariable structure function has an ap- propriate form, as discussed above, frequency re- sponse information for each channel can be used in the analysis of the nominal system in exactly the same way as for the analysis of a conventional feed- back loop in a SISO control system application. How- ever, the multivariable structure function provides additional information about potential interactions and the stability robustness of the closed-loop sys- tem. It is important to note that, for successful applica- tion of the ICAD approach to a two-channel system, the closed-loop bandwidth specification for one chan- nel must not be too similar to the equivalent specifi- cation for the second channel. If this is not the case the problem of the transfer function of one channel depending on the controller of the other channel be- comes a significant obstacle in the processes of anal- ysis and design [2]. 4 Design of the controller for the coupled-tanks system using ICAD For the coupled-tanks system, it may be shown that the multivariable structure function is given by: γ(s) = g12g21 g11g22 = α2 α1+α2 (1 + sA1 α1 )(1 + s A2 α1+α2 ) (23) The expressions for h1(s) and h2(s) for this sys- tem may also be derived directly, from Equations (21) and (22). 4.1 Design requirements The specifications for the closed-loop system were based on equivalent requirements for a SISO version of the coupled-tanks system, for which considerable previous design experience had been accumulated. The requirements for the two-input two-output case were as follows: a) Zero steady-state errors in the liquid levels in both tanks. b) A maximum overshoot of 30 % in liquid levels. c) A damping factor of at least 0.7 which corre- sponds, approximately, to a phase margin of at least 70 degrees for both channels. d) The gain cross-over frequency should be at least 0.05 rad/s for both channels. This value is based on previous experience with the design of PID controllers for the SISO case for control of the liquid level in Tank 2 (with controlled input flow to Tank 1 only), e) For successful application of the ICAD design methodology, it is important to ensure that the polar plots of the multivariable structure func- tions γ(jω), γh1(jω) and γh2(jω) (in terms of the magnitude and phase values at different frequencies over the frequency range of signifi- cance) never approach the point (1, 0). 4.2 An outline of the design process The requirements outlined above provide a basis for design using the ICAD methodology for the linearised plant model, for selected operating conditions. De- sign, in this case, has involved the use of Matlab R© software and has led to continuous and digital con- trollers involving proportional plus integral controller structures for each channel. The design process was carried out for parameter values of the linearised model which correspond to an operating point in the lower half of the depth range in both tanks (H1 = 0.115 m and H2 = 0.071 m). This is a typical operating point for the system un- der open-loop conditions. The values used for the two discharge coefficients are those given in Section 2. The first step in the design process involves es- tablishing that the gain cross-over frequency of the open-loop transmittance of one channel will be sig- nificantly different from the gain cross-over frequency 125 Acta Polytechnica Vol. 52 No. 4/2012 of the other. In this case it was decided, on the basis of physical reasoning, that the gain cross-over fre- quency of Channel 1 should be higher than that for Channel 2. From the design requirements, this latter value should be chosen to be at least 0.05 rad/s, so a value of at least 0.5 rad/s was required for the gain crossover frequency of Channel 1. The next step involves evaluation of the magni- tude and phase of the multivariable structure func- tion over the range of frequencies that are of impor- tance for the intended application. Figure 5 is a typi- cal plot of the multivariable structure function in po- lar form showing the magnitude and phase of γ(jω) for the complete range of relevant frequencies, and it is clear that the resulting plot involves small values of magnitude and does not come close to the (1, 0) point. This is satisfactory for the operating point considered but similar plots should be considered for a range of different operating conditions. Figure 5: Plot of the multivariable structure function in polar form for the coupled-tanks system showing the magnitude and phase of γ(jω) for the complete range of relevant frequencies for one operating point Next, it is necessary to design the controller k2(s) since the requirements in terms of gain cross-over fre- quency for Channel 2 are less demanding than for Channel 1. Equation (19) shows that the equation for Channel 2 involves the transfer function h1(s) and the known multivariable structure function γ(s). The first step is to assume either that h1(s) = 0 or that h1(s) = 1 and design the controller k2(s) ini- tially on that basis [2]. In this application it was as- sumed that h1(s) = 0, but an initial assumption that h1(s) = 1 would have been equally appropriate. The transfer function for g22(s) given in Equation (17) has a magnitude at low frequencies of 1/α2 and at high frequencies the magnitude decreases in an approxi- mately linear fashion at −20 dB per decade. In order to meet the design requirement of zero steady-state closed-loop error this suggests use of a proportional plus integral type of controller of the form: β1 = (1 + β2s) s (24) where β1 and β2 are constants. This controller will produce infinite gain at zero frequency and thus eliminate any steady state error in the closed-loop system for this channel. The choice of parameter values for the controller involves, ini- tially, the selection of a gain factor β1 to give a suit- able gain cross-over frequency which is at least the required minimum of 0.05 rad/s. The integral action is then considered and the frequency ω = 1 β2 is cho- sen to be sufficiently smaller than the gain cross-over frequency to ensure that the overall phase margin is not influenced to any large extent. Application of this procedure gives the following controller: k2(s) = 0.56 (1 + 8.929s) s (25) Through the use of simulation, closed-loop step re- sponses can be examined (usually on the basis of lin- earised models) and further adjustments can be made in the values for these controller parameters if this is judged to be necessary. After obtaining that first approximation to k2(s) an initial single-input single-output design can be carried out for the controller k1(s) on the basis of h2(s), which is now available (from Equation (21)). The procedure followed is essentially the same as for Channel 1 but with the higher value of gain crossover frequency that is required for this channel. The pro- portional plus integral controller resulting initially from this process has the form: k1(s) = 4.676 (1 + 5.988s) s (26) Having found an initial k1(s), this transfer function can then be used to determine h1(s) by substitution into Equation (22). The resulting gain and phase margins must then be checked and adjustments made to k1(s), if necessary. The process may have to be repeated once or twice. Then, using the revised con- troller transfer function for Channel 1, the design can be completed for Channel 2 using a similar it- erative procedure. Final checks must then be made on both channels to compare the gain cross-over fre- quencies with the design specifications and check that the gain and phase margins are all satisfactory. This process also involves re-examination of the Nyquist plots of the multivariable structure functions γ(jω), γh1(jω) and γh2(jω) to ensure that none of them approaches the point (1, 0) and thus establish that the gain and phase margins are valid measures of ro- bustness [2]. 126 Acta Polytechnica Vol. 52 No. 4/2012 Following the application of the above procedures the optimised controller transfer functions were as follows: k1(s) = 5.0 (1 + 6.2s) s (27) k2(s) = 0.56 (1 + 10.0s) s (28) Figure 6: Bode diagram showing magnitude (dB) and phase (deg) for Channel 1 with the compensation provided by the controller transfer function of Equa- tion (27). The gain cross-over frequency is indicated by the vertical line at frequency of about 0.8 rad/s Figure 7: Bode diagram showing magnitude (dB) and phase (deg) for Channel 2 with the compensation provided by the controller transfer function of Equa- tion (28). The gain cross-over frequency is indicated by the vertical line at frequency of about 0.2 rad/s Figures 6 and 7 show the open-loop Bode plots for Channels 1 and 2, respectively, for these optimised controllers. From these Bode plots it may be seen that the gain crossover frequencies for Channel 1 and Channel 2 are approximately 0.8 rad/s and 0.2 rad/s, respectively. The corresponding phase margins are more than the required value of 70 degrees, in both cases. From the gain-crossover frequencies it is clear that the speed of response for Channel 2 is likely to be about four times slower than for Channel 1, which is consistent with the specifications. Discrete equivalents of these continuous con- trollers have been found and an ICAD-based con- trol system has been implemented with a digital con- troller using a general-purpose computer equipped with analogue-to-digital and digital-to-analogue con- verters. However, all experimental results presented in this paper are for the continuous control case where the controllers have been implemented us- ing a small general-purpose electronic analogue com- puter equipped with comparators and switches that can provide limiting integrator action, if required, to avoid integrator saturation. For purposes of comparison, proportional plus in- tegral controllers have also been designed empirically using the Ziegler-Nichols reaction curve method (see, e.g. [13]). This well-known approach to controller design is based on measurements obtained from sim- ple open-loop tests on the plant. Application of this approach gave the following controller transfer func- tions: k1(s) = 6.98 (1 + 3.96s) s (29) k2(s) = 8.9 (1 + 3.3s) s (30) It should be noted that the two controller transfer functions found from the application of the Ziegler- Nichols approach (Equations (29) and (30)) are very much closer in terms of parameter values than the two controllers found using the ICAD approach, as given in Equations (27) and (28). This is because of the requirement within the ICAD methodology that the bandwidth values for the two channels should be significantly different. 5 Results Extensive analysis and simulation studies have been performed using Matlab R© and Simulink R© to inves- tigate the performance of the system, especially in terms of interactions between the two channels and overall robustness of the control systems. In the case of the control systems derived us- ing the ICAD approach the performance of the con- trollers has also been the subject of detailed ex- perimental investigation in the laboratory using the coupled-tanks system hardware. Interactions be- tween the two channels have been investigated both by experiment and through simulation. For the simulation studies the full nonlinear model has been used, with parameter values as given in Section 2. 127 Acta Polytechnica Vol. 52 No. 4/2012 5.1 Simulation results Figure 8 shows typical simulation results for a test in which simultaneous step changes are applied to the reference inputs determining the required levels in the two tanks, using the continuous controllers of Equations (27) and (28). The resulting simulated changes in liquid levels in Tanks 1 and 2 are shown by the upper and lower traces respectively. In the case of Channel 1 the step change of reference imposed is from 199 mm to 228 mm, while for Channel 2 the change is from 165 mm to 198 mm. This test in- volves input flow values for Tank 1 and Tank 2 which both reach their upper limits for this magnitude of demanded level change. It can be seen from these simulation results that, although the operating point considered is signifi- cantly different from the design point, the design re- quirements have been satisfied and also that the re- sponse of Tank 2 is slower than that of Tank 1, as expected. Figure 8: Simulation results found using the non- linear model with the controllers designed using the ICAD approach for a test in which simultaneous step changes in required levels for Tank 1 and Tank 2 are applied at time t = 100 s. The vertical axis represents liquid level (m) while the horizontal axis represents time (s). The level in Tank 1 is represented by the continuous line while the dashed line represents the level in Tank 2 Investigation, through simulation, of interactions between the two channels have involved introducing a step change of the desired level in one channel while maintaining the original set level in the other. The upper set of simulation results presented in Figure 9a shows the level of liquid in Tank 1 fol- lowing the application of a step change of reference for Channel 1 at time t = 100 s, together with the record for the level in Tank 2. In Figure 9b the lower plot shows the liquid level in Tank 2 following the application, at time t = 100 s, of a step change of reference for Channel 2 while the upper trace shows the corresponding level in Tank 1. a) b) Figure 9: a) Simulated responses of levels (m) in Tank 1 (continuous line) and in Tank 2 (dashed line) versus time (s) when the reference level for Channel 1 is changed. The horizontal axis represents time (s). This test involved use of the nonlinear model with controllers designed using the ICAD approach b) Simulated responses of levels (m) in Tank 1 (con- tinuous line) and in Tank 2 (dashed line) when the reference level for Channel 2 is changed. This test involved use of the nonlinear model with controllers designed using the ICAD approach. The horizontal axis represents time (s) These results show that a transient disturbance occurs in the level of Tank 2 when the set level of Channel 1 is changed but that a negligible transient is found in the level of Tank 1 when the set level of Channel 2 is altered by a similar amount. This dif- ference is due to the different bandwidths in the two channels. Results of an additional simulation test are shown in Figure 10. This involves the simultaneous appli- cation of negative step changes of reference for both channels at time t = 100 s. The demanded changes lead, transiently, to a complete cut-off of input flow for both tanks for a period of about 20 s at the time when the reference values are changed, as can be seen from the almost straight-line form of the negative- going responses in that part of the record. 128 Acta Polytechnica Vol. 52 No. 4/2012 Figure 10: Simulation results showing liquid lev- els (m) for Tank 1 (continuous line) and Tank 2 (dashed line) for large negative step changes of refer- ence. The horizontal axis represents time (s) The closed-loop performance of the system with the proportional plus integral controllers designed us- ing the Ziegler-Nichols reaction curve approach (see, e.g. [13]) was investigated through simulation. It was found that for small changes of the reference inputs the two-input two-output system with these controllers behaved very much in accordance with ex- pectations (as shown in Figure 11). Figure 11: Simulation results found using the nonlin- ear model with controllers designed using the Ziegler- Nichols approach for a test involving relatively small changes of reference level. The vertical axis repre- sents liquid level (m) while the horizontal axis repre- sents time (s). The level in Tank 1 is represented by the continuous line while the dashed line represents the level in Tank 2 For large positive changes of reference (similar to those applied in obtaining the results shown in Fi- gure 8 for the ICAD design) the responses for the control system designed using the Ziegler-Nichols ap- proach are found to be much more oscillatory, as shown in Figure 12. This is also the case for the transients found for large negative reference changes, as shown in Figure 13. Figure 12: Simulation results found using the nonlin- ear model with controllers designed using the Ziegler- Nichols approach for a test similar to that of Figure 8. The vertical axis represents liquid level (m) while the horizontal axis represents time (s). The level in Tank 1 is represented by the continuous line while the dashed line represents the level in Tank 2 Figure 13: Simulation results for the nonlinear model with controllers designed using the Ziegler-Nichols approach for a test involving large negative step changes of reference. The vertical axis represents liquid level (m) while the horizontal axis represents time (s). The level in Tank 1 is represented by the continuous line, whereas the dashed line is the level in Tank 2 Responses found for simulated situations involv- ing interactions between the two tanks were also more oscillatory in nature and varied more with operating point than those found using the controllers designed using the ICAD approach. 5.2 Experimental results As implemented using operational amplifiers and the associated passive components, the two controllers corresponded to the transfer functions (as given in Equations (27) and (28)) that resulted from the final optimisation stage of the design process. These are, of course, also the controller transfer functions used in the simulation studies discussed in Section 5.1. 129 Acta Polytechnica Vol. 52 No. 4/2012 Experimental results for the control system, when implemented with these controllers, are shown in Figure 14 for the case involving two simultaneous changes of reference. The results are almost identical in terms of steady state performance to the simulated results of Figure 8 for the same test conditions, and are very similar in terms of the settling time of the transients. As in the simulation results, the inputs both reach their limits during the transients. The main difference observed between the experi- mental results of Figure 14 and the simulation results of Figure 8 is that the transients found experimen- tally (especially for the level in Tank 1) are more os- cillatory than those found through simulation. Sim- ilar findings have been obtained for equivalent tests at other operating points, and this suggests strongly that there are imperfections within the model of the two-tank system. Exactly what the modelling errors might be is not, of course, clear from the information from these closed-loop system tests alone. Although the results shown in Figures 8 and 14 are for one specific operating condition, comparison of experimental and simulation results for a range of different conditions has shown good overall agree- ment. Figure 14: Experimental results for conditions equiv- alent to those of the simulation results of Figure 8. The continuous line shows the measured liquid level (m) in Tank 1 while the dashed line shows the measured level in Tank 2. The horizontal axis repre- sents time (s) The experimental investigation of interactions be- tween channels produced results shown in Figures 15a and 15b, which can be seen to correspond closely to the corresponding simulation results shown in Fig- ures 9a and 9b. Experimental results for a test involving simul- taneous large negative changes of reference value for both channels simultaneously are shown in Figure 16. These results are very similar in character to the sim- ulated results of Figure 10. As in the simulation, the controlled flows for Tank 1 and Tank 2 reach limiting values (zero) during the transient period. a) b) Figure 15: a) Experimental results, equivalent to the simulation results of Figure 9a, involving application of a step change of reference for Channel 1 while the reference input for Channel 2 is held constant. The record for liquid depth (m) in Tank 1 is shown by a continuous line and for Tank 2 by the dashed line. The horizontal axis represents time (s) b) Experimental results, equivalent to the simulation results of Figure 9, involving application of a step change of reference for Channel 2 while the reference for Channel 1 is held constant. The record for liquid depth (m) in Tank 1 is shown by a continuous line and for Tank 2 by the dashed line. The horizontal axis represents time (s) Figure 16: Experimental results showing liquid lev- els (m) for Tank 1 (continuous line) and Tank 2 (dashed line) for large negative step changes of refer- ence. The horizontal axis represents time (s) 130 Acta Polytechnica Vol. 52 No. 4/2012 Figure 17: Experimental results for a test involving the addition of a small volume of water to Tank 1 (continuous trace) and to Tank 2 (dashed line) in turn. The horizontal axis represents time (s) a) b) Figure 18: a) Results of a simulated test involving the addition of a small volume of water to Tank 1 (con- tinuous line). The level in Tank 2 is shown by the dashed line. The horizontal axis represents time (s) b) Results of a simulated test involving the addition of a small volume of water to Tank 2 (dashed line). The level in Tank 1 is shown by the continuous line. The horizontal axis represents time (s) The behavior of the control system when sub- jected to external disturbances is also of great practi- cal importance. Figure 17 shows some typical experi- mental results where external disturbances have been introduced by adding, in as short a period of time as possible, a disturbance in the form of a measured vol- ume of water to each of the tanks in turn, with feed- back control loops applied. The upper plot shows the level in Tank 1 for a reference input of 227 mm, while the lower plot shows the level in Tank 2 for a refer- ence input of 198 mm. The disturbance inputs are applied by the rapid addition of water, from a beaker, to Tank 1 and addition of a similar volume to Tank 2 at about time t = 225 s. The effects of each of these disturbances on each channel are clearly visible in these records. The results show the distinctive actions of the two channels in countering the effects of the distur- bance inputs. The levels for both channels return to their set values after the disturbances, with tran- sients of acceptable magnitude and duration. As with the tests involving changes of reference, it is clear (as would be expected) that the speed of response to dis- turbances is influenced by the choice of bandwidths for the two channels. Similar results have been obtained through simu- lation, but quantitative comparisons are difficult for this type of test because it is hard to reproduce the detailed time-course of the disturbance input within the simulation. Figures 18a and 18b show typical simulation results for disturbance tests which are approximately equivalent to the experiments of Fi- gure 17. The experimental and simulation results are seen to be qualitatively consistent. 5.3 Results of additional simulation-based investigations One interesting practical finding, which has been fully supported by simulation results, is that control of the level in Tank 1 can only be achieved for condi- tions in which the demanded level in Tank 1 is equal to or greater than that in Tank 2. This is understand- able in terms of the physics of the system since Tank 1 has only one outlet (to the second tank), whereas Tank 2 has two outlets (one to the first tank and the second through the drain pipe). If the demanded value of H2 is greater than the demanded level of H1, liquid will flow into Tank 1 from Tank 2 as well as from the input but no liquid will flow out. Since the input flow cannot become negative, satisfactory control of the level in Tank 1 is impossible in these conditions. Typical experimental results are shown in Figure 19 and these demonstrate, in this specific case, that a demanded level of 0.105 m in Tank 1 cannot be achieved in combination with a larger de- manded level (in this case 0.131 m) in Tank 2. What 131 Acta Polytechnica Vol. 52 No. 4/2012 happens in practice is that the final levels in both tanks become equal to the final demanded level in Tank 2. Figure 20 shows results obtained from simu- lation for a very similar set of conditions. Simulation investigations have confirmed that the addition of a drain pipe to the first tank eliminates this problem and would allow independent control of the two levels for any combination of reference values. Figure 19: Levels (m) found in Tank 1 (continu- ous line) and Tank 2 (dashed line) for a demanded change of the reference for Channel 1 from 0.082 m to 0.105 m and for Channel 2 from 0.048 m to 0.131 m. The horizontal axis represents time (s) Figure 20: Simulated results for a test which involves conditions which are very similar to those of the ex- periment of Figure 16. In this case, following the step changes of reference inputs, the demanded level in Tank 2 is again greater (0.131 m) than the demanded level in Tank 1 (105 mm). Here the continuous line again represents the liquid level in Tank 1 and the dashed line the level in Tank 2. The horizontal axis represents time (s) Another area for further investigation through simulation relates to tests of robustness to changes within the plant. These have involved, for exam- ple, the introduction of sudden changes of the cross- sectional area of the outlet drain orifice, or of the cross-sectional area of the orifice responsible for the coupling between Tank 1 and Tank 2. Experimental testing is straightforward in the case of the outlet from Tank 2, for which the drain tap can be used, but investigation of changes of the inter-tank orifice area presents practical difficulties since the variation of cross-sectional area normally requires the removal or insertion of a rubber bung for one of the three orifices in the partition that separates the two tanks. Even partial closing and re-opening of the outlet drain tap is difficult to achieve manu- ally in a precise and repeatable fashion. Simulation can therefore be used to advantage to investigate the performance of the system for this type of change. Figure 21: Results of an experiment involving partial closure and re-opening of the drain tap from Tank 2. The action of closure occurs at about t = 60 s and the re-opening takes place at about t = 150 s. The continuous line shows the level (m) in Tank 1 and the dashed line represents the level (m) in Tank 2. The horizontal axis represents time (s) Figure 22: Results from a simulation involving in- stantaneous changes of the cross-sectional area of the orifice representing the drain tap from Tank 2. Par- tial closure occurs at t = 100 s and the re-opening takes place at t = 300 s. The continuous line shows the level (m) in Tank 1 and the dashed line represents the level (m) in Tank 2 132 Acta Polytechnica Vol. 52 No. 4/2012 Inevitably, the results of simulation tests differ slightly from tests carried out on the real system. Typical experimental results showing the effects of changes of drain tap opening and changing the cross- sectional area of the inter-tank orifice are given in Fi- gure 21. Results from simulation, for instantaneous changes of the cross-sectional area of the orifice rep- resenting the drain-tap and outlet pipe, are shown in Figure 22 and these are broadly similar to the experimental findings of Figure 21. Both the sim- ulation results and the experimental findings confirm that the robustness properties of the two-input two- output control system, as displayed in this test, are satisfactory. Transients for Tank 2 are significantly larger than for Tank 1 as could be expected from the bandwidths of the two channels. 6 Discussion and conclusions The work reported in this paper illustrates the use of the ICAD analysis and design approach for a prac- tical application that involves significant nonlineari- ties in terms both of control input limits and inherent nonlinearity of the plant model. Analysis of the two- input two-output system within the ICAD framework provides helpful insight which can be used in the de- sign and implementation of the control system. Comparisons between simulation and experimen- tal results also provide useful information about the system performance and about limitations of the plant model. A previous journal paper [9] report- ing the application of the ICAD methodology to the same system was concerned with issues of controller parameter tuning and did not consider the response to disturbances or address robustness issues. It can be concluded that the coupled tanks equip- ment provides a useful test-bed for investigation of is- sues of nonlinear system modeling, multivariable con- trol system design and implementation. The avail- ability of a comprehensive nonlinear model of the system, together with linearised representations ap- propriate for control system design, also makes this system suitable for the teaching of practical aspects of multi-input multi-output control system analysis and design using ICAD or other approaches. It is be- lieved that the work reported in this paper could pro- vide the basis for a useful case-study (most probably for use at postgraduate level) on the ICAD methodol- ogy. This could also illustrate the benefits of bringing together simulation and experimental testing within the processes of control system design and implemen- tation. Differences between simulation results and exper- imental results are believed to relate mainly to limita- tions in the representation of the plant and especially the relationship used to describe the output flow from the second tank within the nonlinear model. This as- pect of the coupled-tanks system model has been dis- cussed in previous model validation studies for this system (e.g. [7,10,11]) and is the subject of ongoing investigations. Simulation results show that broadly satisfac- torily results can also be obtained for this plant with proportional plus integral controllers designed empirically using the Ziegler-Nichols reaction curve method. However, results found using that approach have given responses, for the types of tests described in this paper, that tend to be more oscillatory than those found using the ICAD methodology, and indi- cate some issues of closed-loop system robustness to changes of operating point, variation of the magni- tude of reference changes and the magnitude of dis- turbances. Claims that the ICAD approach can provide en- hanced performance compared with other available design methods would be inappropriate on the ba- sis of the limited results presented in this single ap- plication. However, it is believed that the ICAD methodology brings more physical insight to the de- sign process for multi-input multi-output systems, and it must also be remembered that this approach is not restricted to one single form of controller. Acknowledgement This paper is a modified and extended version of a paper [8] presented at EUROSIM 2010 (the 7th EUROSIM Congress on Modelling and Simulation), which was organized by staff of the Department of Computer Science and Engineering, Czech Techni- cal University in Prague (CTU). The Congress took place in Prague during the period 6-10 September, 2010. The author must thank Jin Lin Chiang who, dur- ing undergraduate project work at the University of Glasgow [12], carried out experiments from which some of the results in this paper have been derived. The author also wishes to thank Professor John O’Reilly of the University of Glasgow for many valu- able discussions about ICAD methods. References [1] Wellstead, P. E.: Coupled Tanks Apparatus: Manual. TecQuipment Ltd., UK, 1981. [2] O’Reilly, J., Leithead, W. E.: Multivariable con- trol by ‘individual channel design’. International Journal of Control, 54, 1991, p. 1–46. [3] Leithead, W. E., O’Reilly, J.: Performance is- sues in the individual channel design of 2-input 2-output systems. 1. Structural issues. Interna- tional Journal of Control, 54, 1991, p. 47–82. 133 Acta Polytechnica Vol. 52 No. 4/2012 [4] Leithead, W. E., O’Reilly, J.: Performance is- sues in the Individual Channel Design of 2-input 2-output systems 2. Robustness issues. Interna- tional Journal of Control, 55, 1992, p. 3–47. [5] O’Reilly, J., Leithead, W. E.: Frequency-domain approaches to multivariable feedback control system design: An assessment by individual channel design for 2-input 2-output systems. Control Theory and Advanced Technology, 10, 1995, p. 1 913–1940. [6] Kocijan, J.: Bibliography on Individual Channel Analysis and Design Framework. http://dsc.ijs.si/jus.kocijan/ICAD. (Accessed suc- cessfully 1. 11. 2011). [7] Gong, M., Murray-Smith, D. J: A practical exer- cise in simulation model validation, Mathemati- cal and Computer Modelling of Dynamical Sys- tems, 4, 1998, p. 100–117. [8] Murray-Smith, D. J.: The Individual Channel Analysis and Design method applied to control of a coupled-tanks system: simulation and ex- perimental results. In “Proceedings of the 7th EUROSIM Congress on Modelling and Simula- tion. (Editors: M. Šnorek, Z. Buk,, M. Čepek, and J. Drchal), Prague : Department of Com- puter Science and Engineering, Czech Techni- cal University in Prague (CTU), 2010. (ISBN 9788001045893). [9] Murray-Smith, D. J., Kocijan, J., Gong, M.: A signal convolution method for estimation of con- troller parameter sensitivity functions for tun- ing of feedback control systems by an itera- tive process. Control Eng. Practice, 11, 2003, p. 1 087–1 094. [10] Tan, K. C., Li, Y., Murray-Smith, D. J. ,Shar- man, K. C.: System identification and lineari- sation using genetic algorithms with simulated annealing. In Proc. Conf. on Genetic Algorithms in Eng. Syst.: Innovations and Applications, 12–14 September 1995, (IEE Conference Publi- cation No. 414), London : IEE, 1995, p. 164–189. [11] Gray, G. J., Murray-Smith, D. J., Li, Y., Sharman, K. C., Weinbrenner, T.: Nonlin- ear model structure identification using genetic programming, Control Eng. Practice, 6, 1998, p. 1 341–1 352. [12] Chiang, J. L.: Control System for a Two-Input Two-Output Liquid Level System. Project Re- port 94-179, University of Glasgow, Department of Electronics and Electrical Engineering, 1994. [13] Golten, J., Verwer, A.: Control System Design and Simulation, McGraw-Hill, London, 1991. 134