Next Article in Journal
Modelling Analysis of Channel Assembling in CRNs Based on Priority Scheduling Strategy with Reserved Queue
Previous Article in Journal
Blockchain and Homomorphic Encryption for Data Security and Statistical Privacy
Previous Article in Special Issue
Concurrent Learning-Based Two-Stage Predefined-Time System Identification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Soft Sensors for Industrial Processes Using Multi-Step-Ahead Hankel Dynamic Mode Decomposition with Control

by
Luca Patanè
,
Francesca Sapuppo
* and
Maria Gabriella Xibilia
Department of Engineering, University of Messina, Contrada di Dio, 98158 Messina, Italy
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(15), 3047; https://doi.org/10.3390/electronics13153047
Submission received: 27 May 2024 / Revised: 19 July 2024 / Accepted: 30 July 2024 / Published: 1 August 2024
(This article belongs to the Special Issue Nonlinear System Identification and Soft Sensor Design)

Abstract

:
In this paper, a novel data-driven approach for the development of soft sensors (SSs) for multi-step-ahead prediction of industrial process variables is proposed. This method is based on the recent developments in Koopman operator theory and dynamic mode decomposition (DMD). It is derived from Hankel DMD with control (HDMDc) to deal with highly nonlinear dynamics using augmented linear models, exploiting input and output regressors. The proposed multi-step-ahead HDMDc (MSA-HDMDc) is designed to perform multi-step prediction and capture complex dynamics with a linear approximation for a highly nonlinear system. This enables the construction of SSs capable of estimating the output of a process over a long period of time and/or using the developed SSs for model predictive control purposes. Hyperparameter tuning and model order reduction are specifically designed to perform multi-step-ahead predictions. Two real-world case studies consisting of a sulfur recovery unit and a debutanizer column, which are widely used as benchmarks in the SS field, are used to validate the proposed methodology. Data covering multiple system operating points are used for identification. The proposed MSA-HDMDc outperforms currently adopted methods in the SSs domain, such as autoregressive models with exogenous inputs and finite impulse response models, and proves to be robust to the variability of systems operating points.

1. Introduction

Soft sensors (SSs) are widely exploited for monitoring and controlling industrial processes where real-time estimation of variables is essential. SSs provide mathematical solutions for estimating and predicting hard-to-measure variables (i.e., quality features) using easy-to-measure variables (i.e., quantity features) [1,2].
In automation and control strategies, SSs are applied to address the issues related to delays in measurements due to time-consuming laboratory analysis in the feedback loop. This requires that the designed SSs are capable of multi-step-ahead prediction. Applications are ubiquitous in the process industry: refineries [3], chemical plants [4], power plants [5], food processing [6], polymerization processes [7] or wastewater treatment systems [8]. Industrial processes are often highly nonlinear, suffer from intrinsic dynamic dependencies between input and output variables and may exhibit transients, intermittent phenomena and continuous spectra.
The main approach to identifying such complex nonlinear dynamical systems originates with Poincare’s studies and works on the geometry of subspaces of local linearizations around fixed points, periodic orbits and more general attractors [9]. This methodology has a deep theoretical foundation, such as the Hartman–Grobman theorem, which determines when and where it is possible to approximate a nonlinear system with linear dynamics.
On the one hand, such a geometric perspective enables the application of simple quantitative, locally linear models, such as autoregressive (ARX), principal component regression (PCR) and partial least-square regression (PLSR) models [1], and proper orthogonal decomposition [10], as well as the composition of multiple linear systems as components of more complex modeling techniques [11,12]. In this scenario, the rich linear analytical framework can be used around such operating points and is therefore suitable for linear control strategies. On the other hand, the global analysis remains qualitative and is based on computational analysis, which is not suitable for predicting, estimating and controlling nonlinear systems far from fixed points and periodic orbits. Moreover, due to the complex theoretical environment, data-driven approaches are often used for SSs to support both linear and nonlinear methods [12,13,14,15,16,17]. In this methodological and application scenario, the Koopman operator [18] can provide a theoretical tool for obtaining a global linear representation that is valid for nonlinear systems, even far from fixed points and periodic orbits. A main motivation for the adoption of the Koopman framework is the possibility to simplify the dynamics by the eigenvalue decomposition of the Koopman operator [19] and thus to represent a nonlinear dynamical system globally by an infinite-dimensional linear operator. It uses a Hilbert space of observable functions related to the state of the system to describe the space of all possible measurement state functions. It is linear and its spectral decomposition fully characterizes the behavior of a nonlinear system, without a direct relation to the operating points of the system.
The application of such a powerful tool to industrial problems, namely to obtain a finite-dimensional approximation of the Koopman operator, is a challenge of recent research [20]. Moreover, since the closed form of the Koopman operator is not always obtainable [21], data-driven algorithms are needed. Koopman mode decomposition can be performed using data-driven approaches such as dynamic mode decomposition (DMD) [22].
Applications of DMD can be found in the literature in fluid dynamics [23,24], epidemiology [25], neuroscience [26], plasma physics [27,28], robotics [29], power grid instabilities [30] and renewable energy prediction [31]. DMD represents a method for approximating the Koopman operator that provides a best-fit linear model for one-step-ahead prediction. Such an approximation might not be rich enough to describe nonlinear dynamics. To overcome this limitation and apply DMD to nonlinear industrial processes, it is possible to extend DMD with different strategies based on either nonlinear functions or delayed measurements. Extended DMD [32] and sparse identification nonlinear dynamics (SINDy) [33] belong to the first category. The second category, on which this paper focuses, is based on the use of delayed state variables obtained by the Hankel operator. Such approaches overcome the limitations of standard DMD, which cannot accurately describe systems where the number of variables is smaller than the spectral complexity. Therefore, in Hankel-based DMD, the number of variables is increased by considering time-delayed vectors in addition to the current state vector. There are a few variants of the Hankel approach for the Koopman operator: Hankel DMD (HDMD) [34], high-order DMD (HODMD) [35] and Hankel alternative view of Koopman (HAVOK) [36]. Thanks to the state variable augmentation, these methods are more robust and accurate than classical DMD and are therefore suitable for the identification of nonlinear dynamical systems and offer robust noise filtering [37,38,39].
With the aim of applying DMD-based approaches to industrial processes with exogenous inputs, the DMD with control (DMDc) approach has been proposed in the literature [40]. This is a modified version of DMD that considers both system measurements and exogenous control inputs to identify input–output relationships and the underlying dynamics. Hankel DMD with control (HDMDc) has recently been introduced to handle both time-delayed state variables and control inputs [41,42]. Applications of Koopman theory for quasiperiodically driven systems have also been presented in [43].
In this paper, we propose an extension of the HDMDc approach to multi-step-ahead (MSA) prediction (hereafter referred to as MSA-HDMDc) in the SSs design domain. This solution leverages and exploits the intrinsic HDMDc capability for forecasting [42], operating and, in addition, iterative multi-step-ahead model optimization and output prediction. This makes SSs suitable for the application of model-based online control strategies that are widely used in industrial processes, such as model predictive control (MPC) [44,45,46,47]. To evaluate the potential of the MSA-HDMDc approach in industry, and to test the robustness and reliability of the method in real-world industrial environments considering noise and uncertainty [48,49], two widely used benchmarks in the SS field are considered: the sulfur recovery unit (SRU) [1,12,17,50,51] and the debutanizer column (DC) [52,53]. Multi-step-ahead prediction of the output variables is evaluated on such datasets, and a comparison with currently adopted linear model identification techniques is performed.
The main outcomes of this work are summarized here:
  • The MSA-HDMDc procedure is developed and applied in the field of soft sensors for industrial applications to perform multi-step-ahead prediction;
  • A model order reduction strategy with a two-step optimization is developed to reduce the computational complexity of the identified model;
  • A global linear model for a nonlinear process is identified so that model analysis and control cover multiple working points of the process.
The article is structured as follows. Section 2 describes the theoretical background, the equations and the algorithmic procedure behind the models used. The two industrial case studies, the SRU and the DC, are presented in Section 3Section 4 and Section 5 present the simulation results for the implemented models, including the design and optimization of hyperparameters and model order reduction for both case studies. Comparisons with baseline methods are also reported in this section. Finally, conclusions are drawn in Section 6.

2. Theory Fundamentals

There are different approaches for the data-driven identification of dynamic processes. When dealing with linear models, the widely used model classes in the SS field are the AutoRegressive with eXogenous Input (ARX) and the finite impulse response (FIR) filter [1,2]. They are considered here for comparison with the new MSA-HDMDc, which has been evaluated in industrial applications. This section describes the theoretical and mathematical foundations for both the baseline and the HDMDc models. The algorithmic procedure of the MSA-HDMDc is also presented.

2.1. Baseline Methods

An ARX model set is determined by two polynomials whose degrees are n a and n b , respectively:
A ( z 1 , θ ) = 1 + a 1 z 1 + a 2 z 2 + + a n a z n a B ( z 1 , θ ) = b 0 + b 1 z 1 + b 2 z 2 + + b n b z n b
where z 1 represents the time delay operator and θ is the set of parameters:
θ : = [ a 1 a 2 a n a b 1 b 2 b n b ] T
The acronym ARX can be explained in the model equation form for the calculation of y ( t ) , the predicted output at the time instant t:
y ( t ) = B ( z 1 , θ ) A ( z 1 , θ ) u ( t ) + 1 A ( z 1 , θ ) e ( t )
or equivalently:
A ( z 1 , θ ) y ( t ) = B ( z 1 , θ ) u ( t ) + e ( t )
where e ( t ) is a zero-mean white noise process and u ( t ) is the exogenous input vector.
AR refers to the autoregressive part A ( z 1 , θ ) y ( t ) in the model, while X refers to the exogenous term B ( z 1 , θ ) u ( t ) . The model set is completely determined once the integers n a , n b and the parameter set θ have been specified.
A more general expression including an input/output delay is represented by:
y ( t ) = z n k B ( z 1 , θ ) A ( z 1 , θ ) u ( t ) + 1 A ( z 1 , θ ) e ( t )
where n k is the number of the input–output delay samples.
When an SS is designed to replace the hardware sensors, the output regressors are not always available. In these cases, an infinite-step prediction should be performed, using as output regressors the past estimated values. As an alternative, it is preferred to not involve output regressors in the system dynamics description, and finite impulse response (FIR) models can be adopted. FIR is a special case of Equation (1), with n a = 0 .

2.2. HDMDc Method

The algorithm produces a discrete state-space model, hence the notation for discrete instances, x k , of the continuous time variable, x ( t ) , is used, where x k = x ( k T s ) and  T s is the sampling time of the model. Delay coordinates (i.e., x k 1 , x k 2 , etc.) are also included in the state-space model to account for state delay in the system. This procedure allows the creation of the augmented state space relevant to model nonlinear phenomena, as discussed in Section 1. Therefore, we define a state delay vector as:
x d k = [ x k 1 x k 2 x k q + 1 ] T ,
where q is the number of delay coordinates (including the current time step) of the state, with  x d k R ( q 1 ) n x and n x the number of state variables.
The input delay vector is defined as:
u d k = [ u k 1 u k 2 u k q u + 1 ] T ,
where q u is the number of delay coordinates (including the current time step) of the inputs, with  u d k R ( q u 1 ) n u and n u the number of the exogenous input variables.
The discrete state-space function is defined as:
x k + 1 = A x k + A d x d k + B u k + B d u d k ,
where A R n x × n x is the state matrix, A d R n x × ( q 1 ) n x is the state delay system matrix, B R n x × n u is the input matrix and B d R n x × ( q u 1 ) n u is the delay input matrix. The system output is assumed to be equal to the state, i.e., the output matrix is assumed to be the identity matrix. When dealing with system identification in which only an input/output time series is available, this assumption implies that n x should be chosen as the size of the process output vector. The training time series consists of discrete measurements of the outputs (i.e., y k = x k ) and corresponding inputs (i.e., u k ).
The training data exploring the augmented state space, thanks to the delay shifts, are organized in the following matrices:
X = x q x q + 1 x q + 2 x ( q 1 ) + w
X = x q + 1 x q + 2 x q + 3 x q + w
X d = x q 1 x q x q + 1 x ( q 1 ) + w 1 x 2 x 3 x 4 x w + 1 x 1 x 2 x 3 x w
X d = x q x q + 1 x q + 2 x ( q 1 ) + w x 3 x 4 x 5 x w + 2 x 2 x 3 x 4 x w + 1
Γ = u q u q + 1 u q + 2 u ( q 1 ) + w
Γ d = u q 1 u q + 0 u q + 1 u ( q 1 ) + w 1 u ( q q u ) + 2 u ( q q u ) + 3 u ( q q u ) + 4 x ( q q u ) + w + 1 u ( q q u ) + 1 x ( q q u ) + 2 u ( q q u ) + 3 u ( q q u ) + w
where w represents the time snapshots and is the number of columns in the matrices, X is the matrix X shifted forward by one time step, X d is the matrix with delay states and Γ is the matrix of inputs. Moreover, to incorporate the dynamic effect of control inputs, an extended matrix of the exogenous inputs with time shifts (i.e., Γ d ) is created and included in the model. Equation (8) can now be combined with the matrices in Equations (9)–(14) to produce:
X X d = A X + A d X d + B Γ + B d Γ d
Note that the primary objective of HDMDc is to determine the best-fit model matrices, A, A d , B and B d , given the data in X , X, X d , Γ and Γ d  [40]. Considering the definition of the Hankel matrix, H, for a generic single measurement time series, h k , and applying a d time shift:
H = h d h d + 1 h d + 2 h ( d 1 ) + w h d 1 h d h d + 1 h ( d 2 ) + w h 1 h 2 h 3 h w
we can introduce the synoptic notation:
X H = X X d , X H = X X d , Γ H = Γ Γ d , A H = A A d , B H = B B d
with X H R q n x × w and Γ H R q u n u × w the Hankel matrices for the time series x k and u k , respectively. A H and B H are the transformation matrices for the augmented state and inputs, with  A H R q n x × q n x and B H R q n x × q u n u .
Considering the matrix Ω R ( q n x + q u n u ) × w as the composition of the delayed inputs and outputs, and G as the global transformation matrix described in Equation (18).
Ω = X H Γ H , G = A H B H ,
we obtain:
X H = G Ω
A truncated Singular Value Decomposition (SVD) of the Ω matrix results in the following approximation:
Ω U ˜ p Σ ˜ p V ˜ p T
where the notation ˜ represents rank-p truncation of the corresponding matrix, U ˜ R ( q n x + q u n u ) × p , Σ ˜ R p × p , and  V ˜ R w × p . Then the approximation of G can be computed as:
G X H V ˜ p Σ ˜ p 1 U ˜ p T
For reconstructing the approximate state matrices A ˜ H and B ˜ H , the matrix U ˜ p can be split in two separate components: U ˜ p 1 , related to the state, and U ˜ p 2 , related to the exogenous inputs:
U ˜ p T = U ˜ p 1 T U ˜ p 2 T
where U ˜ p 1 R q n x × p and U ˜ p 2 R q u n u × p .
The complete G matrix can be therefore split in:
G A ¯ H B ¯ H = X H V ˜ p Σ ˜ p 1 U ˜ p 1 T X H V ˜ p Σ ˜ p 1 U ˜ p 2 T
Due to the high dimension of the matrices and to obtain further optimization in the computation of the reconstructed system, a truncated SVD of the X H matrix results in the following approximation:
X H U ^ r Σ ^ r V ^ r T
where the notation ^ represents rank-r truncation, U ^ r R q n x × r , Σ ^ r R r × r , and V ^ r R w × r , and typically we consider r < p . Considering the projection of the operators A ¯ H and B ¯ H on the low-dimensional space we obtain:
A ˜ H = U ^ r T A ¯ H U ^ r = U ^ r T X H V ˜ p Σ ˜ p 1 U ˜ p 1 T U ^ r
B ˜ H = U ^ r T B ¯ H = U ^ r T X H V ˜ p Σ ˜ p 1 U ˜ p 2 T
with A ˜ H R r × r and B ˜ H R r × q u n u The approximated discrete-time system based on the Hankel transformation of the original time series (i.e., x ˜ k H and u k H ) can be therefore represented as:
x ˜ k + 1 H = A ˜ H x ˜ k H + B ˜ H u k H
with x k H = U ^ r x ˜ k H . The original time series x k is then extracted from x k H considering only the rows with index i = n · q + 1 , where n = 0 , 1 , , ( n x 1 ) .
Figure 1 clarifies the HDMDc procedure at a higher level. The input and output measurement data from the historical dataset of an industrial process are fed into the HDMDc block that performs in sequence the Hankel transformation of the input/output variables, the merging of the state and input matrices, which integrates the control signals (DMDc), and then the model identification procedure (DMD). The DMD performs a space transformation, based on the SVD, and returns the reduced estimated state-space system representation used for the output multi-step-ahead prediction.

2.3. MSA-HDMDc Method

In this section, the proposed MSA-HDMDc method is described and the procedure is presented in Algorithm 1. It performs a model optimization and multi-step-ahead prediction of the process output on the basis of a model identified using HDMDc. The model optimization is based on a cost function depending on a combination of key performance indicators (KPIs) such as the mean average percentage error ( M A P E % ) and the coefficient of determination ( R 2 ), adopted for the comparison of the estimated output to the measured target.
Moreover, to compare a set of models to a chosen baseline ( b l ) one, the performance improvement ( P I % ) index is defined for each KPI as:
K P I P I % = ( T K P I b l ) ( T K P I n e w m o d e l ) ( T K P I b l ) %
with T = 0 for M A P E % and T = 1 for R 2 .
As a preliminary step, the optimal delay shifts q o p t and q u o p t , for the state and the input, respectively, should be determined in sequence. The selection is performed by applying the HDMDc algorithm without order reduction, adopting an optimization algorithm (e.g., grid search strategy), and comparing the prediction performances in terms of M A P E % and R 2 , as will be shown in Section 4.
The MSA-HDMDc algorithm is described in the following. The acquired input/output data samples are required to perform the model identification procedure. The training dataset is used to create the X and Γ matrices with the output and inputs, respectively. The state and input augmentation ( X H and Γ H ) are performed by applying the Hankel operator to the original measurements, and the extended state matrix, Ω , is obtained by appending the X H and Γ H matrices. The core SVD algorithm is performed on the Ω and X H matrices. The iteration for model optimization on multi-step-ahead prediction is then performed by determining, in sequence, the optimal reduction for the Ω and X H matrices. The core of the model reduction and reconstruction is performed in the function Reconstruct in Algorithm 2. It performs four operations:
  • The state matrices order truncation (p,r) for model reduction;
  • The determination of the HDMDc operators A ˜ H and B ˜ H as the state space representation of the identified reduced model;
  • The iterative reconstruction of the multi-step-ahead estimated state in the reduced state space ( X ˜ H ), for each selected time horizon within K m a x ;
  • The remapping of the reduced state variables to the original augmented state space ( X ^ H );
  • The extraction of the original state variables from X ^ H , selecting the rows related to the first time shift or each state variable;
  • The evaluation of the M A P E % and R 2 , comparing the model predictions with the target output.
The model reduction is performed in two steps. In the first step, the  Ω matrix is reduced by adopting different order reductions, p p r a n g e , selected on a grid with maximum value p m a x = n x q o p t + n u q u o p t . In this phase, the matrix X H is kept to the full-order r = r m a x = n x q . The optimal reduction order, p o p t , is determined by maximizing the cost function, f p , over the p r a n g e . The cost function consists of a linear combination of the adopted KPIs ( M A P E % and R 2 ) evaluated on the maximum prediction time step, K m a x . The optimization is performed using the validation data contained in the training dataset. In the second phase, the matrix X H is reduced by adopting different order reductions r r r a n g e , selected on a grid with maximum value r m a x = m i n ( n x q o p t , p ) . The reduced matrix, Ω , with order p = p o p t is here considered. The optimal reduction order, r o p t , is determined by maximizing the cost function, f r , over the r r a n g e . The multi-step-ahead prediction can be performed by reconstructing the output dynamics of the optimal identified model.
Algorithm 1 MSA-HDMDc Algorithm
  • Given the training datasets X R n x × w , Γ R n u × w ;
  • X = [ x 1 x 2 x m 1 ]
  • X = [ x 2 x 3 x m ]
  • Γ = [ u 1 u 2 u m 1 ]
  • Set K m a x as the maximum prediction time step value
  • p m a x = n x q o p t + n u q u o p t
  • r m a x = m i n ( n x q o p t , p )
  • Set p r a n g e and r r a n g e with r < p and r < r m a x
  • X H = Hankel(X, q o p t );
  • X H = Hankel( X , q o p t )
  • Γ H = Hankel( Γ , q u o p t )
  • Ω = [ X H T Γ H T ] T
  • Compute the SVD of Ω = U ˜ Σ ˜ V ˜ T
  • Compute the SVD of X H = U ^ Σ ^ V ^ T
  • for  p p r a n g e  do
  •     [ M A P E % ( p , r m a x ) , R ( p , r m a x ) 2 ] = Reconstruct (p, r m a x , K m a x )
  • end for
  • [ M A P E % b l p , R b l p 2 ] = Reconstruct ( p m a x , r m a x , K m a x )
  • f p ( M A P E % ( p , r m a x ) , R ( p , r m a x ) 2 , M A P E % b l p , R b l p 2 ) =
  • M A P E P I % + R P I % 2
  • p o p t = arg max p p r a n g e f p ( p , r m a x )
  • for  r r r a n g e  do
  •     [ M A P E % ( p o p t , r ) , R ( p o p t , r ) 2 ] = Reconstruct ( p o p t , r , K m a x )
  • end for
  • [ M A P E % b l r , R b l r 2 ] = Reconstruct ( p o p t , r m a x , K m a x )
  • f r ( M A P E % ( p o p t , r m a x ) , R ( p o p t , r m a x ) 2 , M A P E % b l r , R b l r 2 ) = M A P E P I % + R P I % 2
  • r o p t = arg max r r r a n g e f r ( p o p t , r )
  • [ M A P E % ( p o p t , r o p t ) , R ( p o p t , r o p t ) 2 ]=
  • Reconstruct ( p o p t , r o p t , K m a x )
Algorithm 2 Reduction and Reconstruction Algorithm
  • function [ M A P E % , R 2 ]   =   Reconstruct (p,r,K)
  • Truncate the SVD matrices at order p: Ω U ˜ p Σ ˜ p V ˜ p T
  • Truncate the SVD matrices at order r: X H U ^ r Σ ^ r V ^ r T
  • Compute the HDMDc operators A ˜ H = U ^ r T X H V ˜ p Σ ˜ p 1 U ˜ p 1 T U ^ r
  • and B ˜ H = U ^ r T X H V ˜ p Σ ˜ p 1 U ˜ p 2 T
  • for  j = 1 , 2 , , w K m a x  do
  •     Get x j H , u j H from the datasets X H and Γ H
  •      x ˜ j H = U ^ r T x j H
  •     for  k = 0 , 1 , , K m a x 1  do
  •          x ˜ j + k + 1 H = A ˜ H x ˜ j + k H + B ˜ H u j + k H
  •          x ^ j + k + 1 H = U ^ x ˜ j + k + 1 H
  •     end for
  •      X ^ H = [ x ^ 1 H x ^ 2 H x ^ w K m a x H ]
  •     Select X ^ H rows for the first time shift for each state variable
  •     Compute M A P E % and R 2
  • end for
  • Return M A P E % and R 2 for the prediction step K

3. Case of Study

This section contains the description of two case studies that are widely used in the field of SSs. Both are from the petrochemical sector and are complex systems with exogenous inputs and strong nonlinearities.

3.1. The Sulfur Recovery Unit

The SRU desulphurization unit considered here is located in a refinery in Sicily (Italy), as described in [54]. SRUs in refineries are used to recover elemental sulfur from gaseous hydrogen sulfide ( H 2 S ) contained in by-product gasses produced during the refining of crude oil and other industrial processes. Since H 2 S is a hazardous environmental pollutant, such a process is of fundamental importance.
The inlets of each SRU line receive two acid gases: MEA gas, rich in H 2 S , and SWS (Sour Water Stripping) gas, rich in H 2 S and ammonia ( N H 3 ). These input gases are combusted in two separate chamber reactors fed by a suitable airflow supply for combustion control. The output gas stream contains residues of H 2 S and sulphur dioxide ( S O 2 ). Normally, the ratio of H 2 S to S O 2 in the tail gas must be maintained, which is specified by a setpoint. An additional secondary airflow is used as an input to improve process control. This variable is the output of a feedback control system and is used to reduce the peak values of H 2 S and S O 2 .
Figure 2 represents a working scheme for an SRU line. The application of SSs is therefore necessary to estimate such concentrations.
The input and output variables available in the SRU historical dataset are listed in Table 1. MSA-HDMDc is used to estimate the H 2 S concentration (i.e., y 1 output).

3.2. The Debutanizer Column

The column is located at ERG Raffineria Mediterranea s.r.l. (ERGMED) in Syracuse, Italy, and is an integral part of a desulphurization and naphtha splitting plant. In the DC, propane (C3) and butane (C4) are extracted from the naphtha stream as overheads [52].
The DC is required to:
  • Ensure sufficient fractionation in the debutanizer;
  • Maximize the C5 content (stabilized gasoline) in the debutanizer overhead (LP gas splitter feed) while complying with the legally prescribed limit;
  • Minimize the C4 (butane) content in the bottom of the debutanizer (feed to the naphtha splitter).
A detailed schematic of the debutanizer column is shown in Figure 3. It includes the following components:
  • E150B heat exchanger;
  • E107AB overhead condenser;
  • E108AB bottom reboiler;
  • P102AB head reflux pump;
  • P103AB feed pump to the LPG splitter;
  • D104 reflux accumulator.
A number of hardware sensors are installed in the plant to monitor product quality. The subset of sensors that are relevant for the described application are listed in Table 2.
The C4 concentration in the bottom flow is estimated as the output of the designed SS. It is not detected on the bottom stream, but at the overheads of the deisopentanization column. The C4 content in the C5 depends solely on the operating conditions of the debutanizer: it can be assumed that the C4 detected in the C5 stream is that which flows from the bottom of the debutanizer. Due to the location of the analyzer, the concentration values are determined with a long delay, which is not exactly known but is constant and probably in the range of 30 min. To improve the control quality of the DC, real-time estimation of both C4 content and C5 content is required. For this purpose, a virtual sensor is needed, which is described in the following sections.

4. Sulfur Recovery Unit Results

In this section, the proposed MSA-HDMDc is applied to the SRU case study and the results are compared with the baseline models. The available dataset consists of about 14,000 samples, with a sampling period of one minute. In addition, 70 % of the dataset is used for model training, 15 % for the validation of the hyperparameters and model order reduction and, finally, the testing is performed on the remaining 15 % of the dataset.

4.1. Model Optimization and Hyperparameter Tuning

An MSA prediction is performed for the output, y 1 . To validate the performance of the procedure, a time horizon of K m a x = 30 steps is selected. To better show the effect of the MSA prediction for different time horizons, values in the interval K { 1 , 5 , 10 , 15 , 20 , 25 , 30 } are considered.

4.1.1. Baseline Models

Two linear models were considered: ARX and FIR. The optimal model order was selected based on the minimum description length (MDL) criterion [55]. In particular, the ARX structure was identified to have eight common poles ( n a = 8 ), two zeros ( n b = 3 ), and no delay for all input variables ( n k = 0 ), while the FIR order parameters resulted to be n b = 10 and n k = 0 .

4.1.2. MSA-HDMDc Model

In a preliminary phase of the iterative MSA-HDMDc procedure, a parametric study was performed based on model performance in terms of M A P E % and R 2 . Such a procedure allowed the definition of the hyperparameters related to the delay shifts q and q u applied to the state and inputs, respectively.
The first step was to determine the optimal state variable delay shift, q. A grid search strategy was applied and, for each q that lies between q = 20 and q = 60 with a step of 10, the full-rank HDMDc model was identified. The estimated output reconstructed at the maximum prediction step (i.e., 30 steps) was compared with the measured output, y 1 . For statistical analysis, the validation dataset was divided into 20 subsets of 100 samples. M A P E % and R 2 were evaluated for each subset, and the corresponding distribution was determined. In particular, the median value of M A P E % was used to select the optimal hyperparameter, q.
Table 3 shows the mean value of M A P E % over the 20 trials considering a 30-step-ahead prediction for the selected q { 20 , 30 , 40 , 50 , 60 } . The best performing model corresponds to q o p t = 40 , as shown by the reported P I % . In a second step, assuming that the optimal value q o p t = 40 is fixed, a further parametric optimization was performed by varying the Hankel shift, q u , applied to the exogenous control inputs in the Γ H matrix. Since q u q o p t , for the system to be causal, the model was identified for each q u belonging to the set q u { 10 , 15 , 30 , 40 } . The estimated output, reconstructed at each considered prediction step, was compared with the measured output, y 1 . Considering the model with q u = q o p t = 40 , corresponding to the maximum allowable value, as the baseline one, the barplot in Figure 4 shows how the system performance changes ( P I % ) , at different prediction steps, by decreasing the value of q u . It can be noticed how the reduction of the Hankel shift, q u , applied to the input variables causes negative P I % and, thus, the decaying of the model performance, both in terms of M A P E % and R 2 for prediction steps higher than 5. This led to the selection of q u o p t = q o p t = 40 as the optimal number of Hankel shifts for inputs and state variables.

4.2. Model Order Reduction

To evaluate the importance of the model reduction phase, as proposed in MSA-HDMDc, the results obtained during the model order optimization are shown for different prediction horizons. To better assess the performance of the model, both M A P E % and R 2 are shown.
According to Algorithm 1, the analysis was performed considering two optimization phases, first for p and then for r, each representing the SVD truncation for the matrices Ω (Equation (20)) and X H (Equation (24)). In a first step, the Ω matrix was truncated from the full-order p m a x = n x q o p t + n u q u o p t = 240 (with q o p t = 40 , n x = 1 , q u o p t = 40 , and n u = 5 ) to the reduced order p. Models with different p-reductions from the set p r a n g e { 201 , 202 , 210 , 220 , 240 } and with the full-order matrix X H with order r m a x = m i n ( n x q o p t , p ) = 40 are shown in Figure 5, where the indices M A P E % and R 2 are given for different prediction steps. The global performance of HDMDc on the validation dataset was compared with the performance of the baseline methods (i.e., A R X and F I R ). It can be noticed that the MSA-HDMDc model outperforms the FIR model for all prediction steps and the ARX model for prediction steps greater than or equal to 5, regardless of the model order, p. The comparison of the MSA-HDMDc model for the different p also shows that the reduction of the Ω matrix to order p has a positive effect on the performance of the model for prediction steps greater than 5, and the optimal configuration is achieved for order p = 202 .
In a second step, while maintaining the optimal Ω order p o p t = 202 , the X H matrix was truncated from the full-order r m a x = m i n ( n x q o p t , p o p t ) = 40 to the reduced order, r. Models with different r reductions belonging to the set r { 18 , 20 , 23 , 25 , 30 , 35 } were identified. The global performance comparison on the validation dataset is shown in Figure 6, where the P I % is given for both M A P E % and R 2 at different prediction steps. The reference model for the P I % is MSA-HDMDc with p o p t = 202 and X H full-order r m a x = 40 . It can be noticed how the MSA-HDMDc model with a reduction of X H to order r = 25 outperforms the full-order model for prediction steps greater than 5. These results confirm that the optimized MSA-HDMDc order reduction allows identification of the dominant dynamics and thus introduces robustness and noise rejection features to the reduced model. These properties thus improve the long-term prediction performance compared with the full-order system. Finally, the optimal order for the MSA-HDMDc model was determined to be p o p t = 202 and r o p t = 25 .

4.3. Model Comparisons and Discussion

In this section, the performance of the MSA-HDMDc reduced-order model is evaluated. In particular, the results for the maximum step prediction (i.e., 30 steps), are here presented. The regression plots on the test dataset, for the baseline and the optimal MSA-HDMDc models, are reported in Figure 7. The ARX model regression plot presents a slope of 0.41 and a bias of 0.13 ; the global performance over the test dataset is not considered acceptable. The FIR regression plot presents a slope of 0.76 with a bias of 0.04 . MSA-HDMDc outperforms both the baseline models with a slope of 0.77 and a bias of 0.068 .
The time plot in Figure 8 shows a comparison between the measured output, y 1 , and the 30-step-ahead predicted output for the baseline and the MSA-HDMDc models for a subset of the test dataset.
As described above, the main objective of the SRU is to remove H 2 S from the gas flow. Therefore, the estimation of the output peaks is of greatest interest for the designed model. Figure 8 shows the prediction performance of the proposed model, which outperforms the results obtained with the baseline approaches, especially with respect to the peak events.
Table 4 and Table 5 reports the M A P E % and R 2 for the test dataset considering different prediction steps for the considered models. Both KPIs are in agreement and show the superiority of the MSA-HDMDc model for large prediction horizons.
As mentioned in Section 1, the strength of the Koopman operator, which forms the basis of MSA-HDMDc, lies in the identification of a global state space model that is valid, even far from specific working points and/or attractors. It differs from standard linear models that either exploit linearization around specific working points or extend the linear approximation to the entire domain. With this in mind, the exogenous inputs in the entire test interval were clustered using the k-means algorithm with squared Euclidean distance, which is commonly used in pattern recognition [56], classification and predictive modeling [57]. Such a method aims to identify different operating points contained in the input dynamics. To select the optimal number of clusters, the silhouette score distribution [58] was used, obtaining three distinct clusters. The results of such a procedure allowed us to divide the test time series into sub-intervals, each of which is associated with a cluster identifying a working point.
The results of the clustering are shown in Figure 9. The first panel shows the time evolution of the exogenous inputs as named and described in Table 1. The second field shows the different operating points, defined as clusters, identified by the k-means algorithm in the analyzed time window. The third field contains the M A P E % time evolution, which was analyzed in time batches of 100 min for a 30-step-ahead prediction. It can be observed that, in the first and last time interval, belonging to C l u s t e r 1 , which is representative of the majority of the dataset, all three models predict the output with good performance. For data belonging to the second and third clusters, only MSA-HDMDc guarantees a performance similar to those obtained in the previous cluster. This confirms the suitability of MSA-HDMDc to identify global models.

5. Debutanizer Column Results

In this section, the proposed MSA-HDMDc is applied to the DC case study and the results are compared with the baseline models. The available dataset consists of 4 months of data, i.e., March, May, July and September 2004, with a sampling period of 6 min. Here, 50 % of the dataset (March and May) is used for model training, 25 % (July) for the validation of the hyperparameters and model order reduction and, finally, the testing is performed on the final 25 % of the dataset (September).

5.1. Model Optimization and Hyperparameter Tuning

An MSA prediction on the y output is performed. To validate the performance of the procedure, a time horizon of K m a x = 20 steps corresponding to 120 min is selected. To better show the effect of MSA prediction for different time horizons, values in the interval K { 2 , 5 , 10 , 20 } steps are considered.

5.1.1. Baseline Models

Two linear models were considered: ARX and FIR. The optimal model order was selected based on the minimum description length (MDL) criterion [55]. In particular, the ARX structure was identified to have three common poles ( n a = 3 ), eight zeros ( n b = 8 ) and no delay for all input variables ( n k = 0 ), while the FIR order parameters resulted to be n b = 8 and n k = 0 .

5.1.2. MSA-HDMDc Model

In a preliminary stage of the iterative MSA-HDMDc method, the HDMDc algorithm was used without order reduction. A grid search strategy was applied to find the optimal q in the range from q = 10 to q = 30 . The estimated output reconstructed for the prediction step, K m a x , was compared with the measured output, y. For the statistical analysis, the validation dataset was divided into 70 subsets of 100 samples. M A P E % and R 2 were evaluated for each subset and the corresponding distribution was determined. In particular, the median value of M A P E % was taken into account when selecting the optimal q hyperparameters.
Table 6 shows the mean value over the 70 trials of the M A P E % , considering a 20-steps-ahead prediction. The considered state delay shifts are q { 10 , 12 , 15 , 17 , 20 , 30 } . The best-performing model corresponds to q o p t = 12 , as shown by the P I % reported.
As a second step, considering the optimal value q o p t = 12 , further parametric optimization was carried out by varying the Hankel shift, q u , applied to the exogenous control inputs in the Γ H matrix. Being q u q o p t , for the system to be causal, the model was identified for each q u belonging to the set q u { 5 , 10 , 12 } . The estimated output, reconstructed at each considered prediction step, was compared with the measured output, y. It was found that the reduction of the Hankel shift, q u , applied to the input variables causes decaying of the model performance, both in terms of M A P E % and R 2 . This led to the selection of q u o p t = q o p t = 12 as the optimal number of Hankel shifts for inputs and state variables.

5.2. Model Order Reduction

In a first step, the Ω matrix was truncated from the full-order p m a x = n x q o p t + n u q u o p t = 84 (where q o p t = 12 , n x = 1 , q u o p t = 12 and n u = 6 ) to the reduced order p. Models with different p reductions in the set p r a n g e { 65 , 66 , 70 , 84 } and with the full-order matrix X H with r m a x = m i n ( n x q o p t , p ) = 12 are shown in Figure 10, where the indices M A P E % and R 2 are given for different prediction steps. The global performance on the validation dataset was compared with the baseline methods (i.e., A R X and F I R ). It can be seen that the MSA-HDMDc model outperforms both FIR and ARX for all prediction steps, regardless of the model order, p. When comparing MSA-HDMDc for the different p, it is also noted that reducing the Ω matrix to order p leaves the performance unchanged, as it decreases from p = 84 to p = 66 , but starts to deteriorate at p = 65 . This leads to the conclusion that p o p t = 66 holds for all prediction steps.
In a second step, while maintaining the optimal Ω order p o p t = 66 , the X H matrix was truncated from the full-order r m a x = m i n ( n x q o p t , p o p t ) = 12 to the reduced order r. Models with different r reductions belonging to the set r { 4 , 5 , 8 , 12 } were identified. The global performance comparison on the validation dataset is shown in Figure 11, where the P I % is given with respect to MSA-HDMDc, with p o p t = 66 and r m a x = 40 for both M A P E % and R 2 at different prediction steps. It can be seen that the MSA-HDMDc model with a reduction of X H to order r = 5 outperforms the full-order model for all selected prediction steps. These results confirm that the optimized MSA-HDMDc order reduction allows to identify the dominant dynamics and thus introduces robustness and noise rejection features to the reduced model. These properties thus improve the long-term prediction performance compared with the full-order system. Finally, the optimal order for the MSA-HDMDc model was determined to be p o p t = 66 and r o p t = 5 .

5.3. Model Comparison and Discussion

In this section, the performance of the reduced-order MSA-HDMDc model is further evaluated in terms of the predicted time series and robustness to the variability of the operating points. The time plot in Figure 12 shows a comparison between the measured output, y, and the 5- and 10-step-ahead predicted outputs for the baseline and MSA-HDMDc models for a subset of the test dataset. It shows that the prediction performance of the proposed model outperforms the results obtained with the baseline approaches, especially for the 10-step-ahead prediction, and confirms that the difference between the models becomes more pronounced at longer prediction intervals.
Table 7 and Table 8 reports the M A P E % and R 2 for the test dataset considering different prediction steps for the considered models. Both KPIs are in agreement and show the superiority of the MSA-HDMDc model for all the prediction horizons.
The analysis of the model performance was performed by identifying different operating points through exogenous inputs clustering. The k-means algorithm [56,57] was applied with squared Euclidean distance. The optimal number of clusters was found to be two by using the silhouette score distribution [58]. The clustering results are reported in Figure 13. In the first panel, the time evolution of the normalized exogenous inputs, as named and described in Table 2, is shown. In the second panel, the clusters identified by the k-means algorithm are reported in the analyzed time window showing the operating points with different colors. The third panel contains the M A P E % time evolution computed in time batches of 100 samples for a 5-step-ahead prediction. It can be seen that, for data belonging to the operating point labeled as Cluster 1 between 1400 and 2000 samples, only the MSA-HDMDc model maintains consistent performance across different system operating points. In contrast, the performance of the ARX and FIR models degrades significantly and shows high variability with changes in operating conditions. This highlights the sensitivity of the ARX and FIR models to variations in operating points and confirms the suitability of MSA-HDMDc for identifying global models that provide robust performance under different operating conditions.

6. Conclusions

In this work, we have proposed the MSA-HDMDc approach for the development of SSs in industrial applications. Two well-established benchmarks based on the SRU and DC data were considered.
The MSA-HDMDc approach focused on optimizing the identified HDMDc state-space model for multiple-step-ahead predictions. It employed the Hankel operator for state-space dimension augmentation to address nonlinear dynamics effectively through a linear state-space model.
As a preliminary step, the augmentation delay time shifts for input and state variables were optimized through statistical analyses, using the M A P E % and R 2 as key performance indexes. As a second step, model order reduction was investigated to find the minimum number of Koopman space observables that better describe the nonlinear system dynamics while keeping the computational burden reasonable, aiming to obtain models that can be integrated into model-based control strategies. An iterative procedure was proposed to select the optimal order reduction, maximizing a fitness function based on performance improvements relative to full-order models. The optimal reduced MSA-HDMDc model structure was then compared for different step-ahead predictions with baseline linear identification methods (i.e., ARX and FIR), which are widely used in the SS field.
The MSA-HDMDc model outperformed, for both the SRU and the DC case studies, the baseline ARX and FIR models for high prediction steps. The model exhibited reduced sensitivity to different operational conditions compared with the baseline models. This was demonstrated through a clustering procedure applied to the input time series data, which identified and classified various operative states as clusters. The performance of the analyzed models was monitored over time and associated with these distinct states. The MSA-HDMDc model consistently maintained its performance across various conditions, whereas the ARX and FIR models showed sensitivity to these variations, resulting in decreased performance for certain clusters. This is in line with the Koopman theory that allows the identification of a global model for the system able to properly work on multiple operating points. This feature, along with the model optimization due to the SVD-based model order reduction, makes the proposed MSA-HDMDc suitable for MPC-based applications, enhancing the reliability and effectiveness of the control system in industrial applications.
The capability of the proposed methodology to identify a global model that functions across different operating points encourages further exploration of its transferability. Applying the identified model to different process lines could address the issue of data scarcity. Additionally, given that SRU and DC data exhibit strong nonlinearity and dynamic variations in operating points and spectral content, future efforts will focus on integrating MSA-HDMDc with multi-resolution DMD [59] to incorporate slow feature analysis.

Author Contributions

Conceptualization, L.P. and M.G.X.; methodology, L.P. and M.G.X.; software, F.S.; validation, L.P. and M.G.X.; formal analysis, L.P. and M.G.X.; investigation, L.P. and M.G.X.; resources, L.P. and M.G.X.; data curation, L.P. and M.G.X.; writing—original draft preparation, F.S.; writing—review and editing, L.P. and M.G.X.; visualization, F.S.; supervision, L.P. and M.G.X.; project administration, L.P. and M.G.X.; funding acquisition, L.P. and M.G.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Progetto Green SENSing systems based on Bacterial Cellulose (SENS-BC), Italian Ministry of University and Research, CUP J53D23003460006.

Data Availability Statement

Dataset available on request from the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
DMDDynamic mode decomposition
HDMDcHankel dynamic mode decomposition with control
MSAMulti-step-ahead
SVDSingular Value Decomposition
MPCModel predictive control
SRUSulfur recovery unit
DCDebutanizer column

References

  1. Fortuna, L.; Graziani, S.; Rizzo, A.; Xibilia, M. Soft Sensors for Monitoring and Control of Industrial Processes; Springer: London, UK, 2007. [Google Scholar] [CrossRef]
  2. Kadlec, P.; Gabrys, B.; Strandt, S. Data-driven Soft Sensors in the process industry. Comput. Chem. Eng. 2009, 33, 795–814. [Google Scholar] [CrossRef]
  3. Pani, A.K.; Amin, K.G.; Mohanta, H.K. Soft sensing of product quality in the debutanizer column with principal component analysis and feed-forward artificial neural network. Alex. Eng. J. 2016, 55, 1667–1674. [Google Scholar] [CrossRef]
  4. Graziani, S.; Xibilia, M.G. Deep Structures for a Reformer Unit Soft Sensor. In Proceedings of the 2018 IEEE 16th International Conference on Industrial Informatics (INDIN), Porto, Portugal, 18–20 July 2018; pp. 927–932. [Google Scholar] [CrossRef]
  5. Sujatha, K.; Bhavani, N.; Cao, S.Q.; Ram Kumar, K. Soft Sensor for Flame Temperature Measurement and IoT based Monitoring in Power Plants. Mater. Today Proc. 2018, 5, 10755–10762. [Google Scholar] [CrossRef]
  6. Zhu, X.; Rehman, K.U.; Wang, B.; Shahzad, M. Modern Soft-Sensing Modeling Methods for Fermentation Processes. Sensors 2020, 20, 1771. [Google Scholar] [CrossRef]
  7. Zhu, C.; Zhang, J. Developing Soft Sensors for Polymer Melt Index in an Industrial Polymerization Process Using Deep Belief Networks. Int. J. Autom. Comput. 2020, 17, 44–54. [Google Scholar] [CrossRef]
  8. Pisa, I.; Santín, I.; Vicario, J.; Morell, A.; Vilanova, R. ANN-Based Soft Sensor to Predict Effluent Violations in Wastewater Treatment Plants. Sensors 2019, 19, 1280. [Google Scholar] [CrossRef]
  9. Guckenheimer, J.; Holmes, P. Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields; Springer: New York, NY, USA, 1983; Volume 42. [Google Scholar] [CrossRef]
  10. Babaei Pourkargar, D.; Armaou, A. Control of spatially distributed processes with unknown transport-reaction parameters via two layer system adaptations. AIChE J. 2015, 61, 2497–2507. [Google Scholar] [CrossRef]
  11. Souza, F.; Araujo, R. Online mixture of univariate linear regression models for adaptive soft sensors. IEEE Trans. Ind. Inform. 2014, 10, 937–945. [Google Scholar] [CrossRef]
  12. Chen, X.; Mao, Z.; Jia, R.; Zhang, S. Ensemble regularized local finite impulse response models and soft sensor application in nonlinear dynamic industrial processes. Appl. Soft Comput. 2019, 85, 105806. [Google Scholar] [CrossRef]
  13. Liu, J.; He, J.; Tang, Z.; Xie, Y.; Gui, W.; Ma, T.; Jahanshahi, H.; Aly, A.A. Frame-Dilated Convolutional Fusion Network and GRU-Based Self-Attention Dual-Channel Network for Soft-Sensor Modeling of Industrial Process Quality Indexes. IEEE Trans. Syst. Man Cybern. Syst. 2022, 52, 5989–6002. [Google Scholar] [CrossRef]
  14. Xie, S.; Xie, Y.; Li, F.; Yang, C.; Gui, W. Optimal Setting and Control for Iron Removal Process Based on Adaptive Neural Network Soft-Sensor. IEEE Trans. Syst. Man Cybern. Syst. 2020, 50, 2408–2420. [Google Scholar] [CrossRef]
  15. Xie, Y.; Wang, J.; Xie, S.; Chen, X. Adversarial Training-Based Deep Layer-Wise Probabilistic Network for Enhancing Soft Sensor Modeling of Industrial Processes. IEEE Trans. Syst. Man Cybern. Syst. 2023, 54, 972–984. [Google Scholar] [CrossRef]
  16. Dias, T.; Oliveira, R.; Saraiva, P.M.; Reis, M.S. Linear and Non-Linear Soft Sensors for Predicting the Research Octane Number (RON) through Integrated Synchronization, Resolution Selection and Modelling. Sensors 2022, 22, 3734. [Google Scholar] [CrossRef]
  17. Patanè, L.; Xibilia, M.G. Echo-state networks for soft sensor design in an SRU process. Inf. Sci. 2021, 566, 195–214. [Google Scholar] [CrossRef]
  18. Koopman, B.O. Hamiltonian Systems and Transformation in Hilbert Space. Proc. Natl. Acad. Sci. USA 1931, 17, 315–318. [Google Scholar] [CrossRef]
  19. Mezić, I. Spectral Properties of Dynamical Systems, Model Reduction and Decompositions. Nonlinear Dyn. 2005, 41, 309–325. [Google Scholar] [CrossRef]
  20. Brunton, S.L.; Budišić, M.; Kaiser, E.; Kutz, J.N. Modern Koopman Theory for Dynamical Systems. SIAM Rev. 2022, 64, 229–340. [Google Scholar] [CrossRef]
  21. Brunton, S.L.; Brunton, B.W.; Proctor, J.L.; Kutz, J.N. Koopman invariant subspaces and finite linear representations of nonlinear dynamical systems for control. PLoS ONE 2015, 11, e0150171. [Google Scholar] [CrossRef]
  22. Brunton, S.L.; Kutz, J.N. Data-Driven Science and Engineering; Cambridge University Press: Cambridge, UK, 2022. [Google Scholar] [CrossRef]
  23. Song, G.; Alizard, F.; Robinet, J.C.; Gloerfelt, X. Global and Koopman modes analysis of sound generation in mixing layers. Phys. Fluids 2013, 25, 124101. [Google Scholar] [CrossRef]
  24. Mezić, I. Analysis of Fluid Flows via Spectral Properties of the Koopman Operator. Annu. Rev. Fluid Mech. 2013, 45, 357–378. [Google Scholar] [CrossRef]
  25. Proctor, J.L.; Eckhoff, P.A. Discovering dynamic patterns from infectious disease data using dynamic mode decomposition. Int. Health 2015, 7, 139–145. [Google Scholar] [CrossRef] [PubMed]
  26. Alfatlawi, M.; Srivastava, V. An Incremental Approach to Online Dynamic Mode Decomposition for Time-Varying Systems with Applications to EEG Data Modeling. arXiv 2019, arXiv:1908.01047. [Google Scholar] [CrossRef]
  27. Taylor, R.; Kutz, J.N.; Morgan, K.; Nelson, B.A. Dynamic mode decomposition for plasma diagnostics and validation. Rev. Sci. Instruments 2018, 89, 053501. [Google Scholar] [CrossRef] [PubMed]
  28. Kaptanoglu, A.A.; Morgan, K.D.; Hansen, C.J.; Brunton, S.L. Characterizing magnetized plasmas with dynamic mode decomposition. Phys. Plasmas 2020, 27, 032108. [Google Scholar] [CrossRef]
  29. Bruder, D.; Gillespie, B.; Remy, C.D.; Vasudevan, R. Modeling and Control of Soft Robots Using the Koopman Operator and Model Predictive Control. arXiv 2019, arXiv:1902.02827. [Google Scholar]
  30. Susuki, Y.; Mezic, I. Nonlinear Koopman Modes and a Precursor to Power System Swing Instabilities. IEEE Trans. Power Syst. 2012, 27, 1182–1191. [Google Scholar] [CrossRef]
  31. Jones, C.; Utyuzhnikov, S. Application of higher order dynamic mode decomposition to modal analysis and prediction of power systems with renewable sources of energy. Int. J. Electr. Power Energy Syst. 2022, 138, 107925. [Google Scholar] [CrossRef]
  32. Williams, M.O.; Kevrekidis, I.G.; Rowley, C.W. A Data–Driven Approximation of the Koopman Operator: Extending Dynamic Mode Decomposition. J. Nonlinear Sci. 2015, 25, 1307–1346. [Google Scholar] [CrossRef]
  33. Brunton, S.L.; Proctor, J.L.; Kutz, J.N. Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proc. Natl. Acad. Sci. USA 2016, 113, 3932–3937. [Google Scholar] [CrossRef]
  34. Arbabi, H.; Mezić, I. Ergodic theory, Dynamic Mode Decomposition and Computation of Spectral Properties of the Koopman operator. SIAM J. Appl. Dyn. Syst. 2016, 16, 2096–2126. [Google Scholar] [CrossRef]
  35. Clainche, S.L.; Vega, J.M. Higher Order Dynamic Mode Decomposition. SIAM J. Appl. Dyn. Syst. 2017, 16, 882–925. [Google Scholar] [CrossRef]
  36. Brunton, S.L.; Brunton, B.W.; Proctor, J.L.; Kaiser, E.; Kutz, J.N. Chaos as an intermittently forced linear system. Nat. Commun. 2017, 8, 19. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, Z.; Brunton, S.L.; Revzen, S. Challenges in dynamic mode decomposition. J. R. Soc. Interface 2021, 18, 20210686. [Google Scholar] [CrossRef] [PubMed]
  38. Clainche, S.L.; Han, Z.H.; Ferrer, E. An alternative method to study cross-flow instabilities based on high order dynamic mode decomposition. Phys. Fluids 2019, 31, 094101. [Google Scholar] [CrossRef]
  39. Clainche, S.L. Prediction of the Optimal Vortex in Synthetic Jets. Energies 2019, 12, 1635. [Google Scholar] [CrossRef]
  40. Proctor, J.L.; Brunton, S.L.; Kutz, J.N. Dynamic Mode Decomposition with Control. SIAM J. Appl. Dyn. Syst. 2016, 15, 142–161. [Google Scholar] [CrossRef]
  41. Mustavee, S.; Agarwal, S.; Enyioha, C.; Das, S. A linear dynamical perspective on epidemiology: Interplay between early COVID-19 outbreak and human mobility. Nonlinear Dyn. 2022, 109, 1233–1252. [Google Scholar] [CrossRef] [PubMed]
  42. Shabab, K.R.; Mustavee, S.; Agarwal, S.; Zaki, M.H.; Das, S. Exploring DMD-Type Algorithms for Modeling Signalised Intersections. arXiv 2021, arXiv:2107.06369v1. [Google Scholar]
  43. Das, S.; Mustavee, S.; Agarwal, S.; Hasan, S. Koopman-Theoretic Modeling of Quasiperiodically Driven Systems: Example of Signalized Traffic Corridor. IEEE Trans. Syst. Man Cybern. Syst. 2023, 53, 4466–4476. [Google Scholar] [CrossRef]
  44. Wolfram, D.; Meurer, T. DMD-Based Model Predictive Control for a Coupled PDE-ODE System. IFAC-PapersOnLine 2023, 56, 4258–4263. [Google Scholar] [CrossRef]
  45. Liu, J.; Zhang, X.; Xu, X.; Xiong, Q. Receding Horizon Actor–Critic Learning Control for Nonlinear Time-Delay Systems with Unknown Dynamics. IEEE Trans. Syst. Man Cybern. Syst. 2023, 53, 4980–4993. [Google Scholar] [CrossRef]
  46. Zhang, X.; Liu, J.; Xu, X.; Yu, S.; Chen, H. Robust Learning-Based Predictive Control for Discrete-Time Nonlinear Systems with Unknown Dynamics and State Constraints. IEEE Trans. Syst. Man Cybern. Syst. 2022, 52, 7314–7327. [Google Scholar] [CrossRef]
  47. Narasingam, A.; Kwon, J.S.I. Development of local dynamic mode decomposition with control: Application to model predictive control of hydraulic fracturing. Comput. Chem. Eng. 2017, 106, 501–511. [Google Scholar] [CrossRef]
  48. Lin, S.; Zhang, M.; Cheng, X.; Shi, L.; Gamba, P.; Wang, H. Dynamic Low-Rank and Sparse Priors Constrained Deep Autoencoders for Hyperspectral Anomaly Detection. IEEE Trans. Instrum. Meas. 2024, 73, 1–18. [Google Scholar] [CrossRef]
  49. Zhang, C.; Li, G.; Lei, R.; Du, S.; Zhang, X.; Zheng, H.; Wu, Z. Deep Feature Aggregation Network for Hyperspectral Remote Sensing Image Classification. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2020, 13, 5314–5325. [Google Scholar] [CrossRef]
  50. Yuan, X.; Wang, Y.; Yang, C.; Gui, W. Stacked isomorphic autoencoder based soft analyzer and its application to sulfur recovery unit. Inf. Sci. 2020, 534, 72–84. [Google Scholar] [CrossRef]
  51. Bidar, B.; Shahraki, F.; Sadeghi, J.; Khalilipour, M.M. Soft Sensor Modeling Based on Multi-State-Dependent Parameter Models and Application for Quality Monitoring in Industrial Sulfur Recovery Process. IEEE Sens. J. 2018, 18, 4583–4591. [Google Scholar] [CrossRef]
  52. Fortuna, L.; Graziani, S.; Xibilia, M. Soft sensors for product quality monitoring in debutanizer distillation columns. Control Eng. Pract. 2005, 13, 499–508. [Google Scholar] [CrossRef]
  53. Mou, T.; Zou, Y.; Li, S. Enhancing graph convolutional network of knowledge-based co-evolution for industrial process key variable prediction. Control Theory Appl. Lilun Yu Yinyong 2024, 41, 416. [Google Scholar]
  54. Fortuna, L.; Rizzo, A.; Sinatra, M.; Xibilia, M. Soft analyzers for a sulfur recovery unit. Control Eng. Pract. 2003, 11, 1491–1500. [Google Scholar] [CrossRef]
  55. Ljung, L. System Identification: Theory for the User; Prentice-Hall, Inc.: Upper Saddle River, NJ, USA, 1986. [Google Scholar]
  56. Sahbudin, M.A.B.; Scarpa, M.; Serrano, S. MongoDB Clustering Using K-means for Real-Time Song Recognition. In Proceedings of the 2019 International Conference on Computing, Networking and Communications (ICNC), Honolulu, HI, USA, 18–21 February 2019; IEEE: Piscataway, NJ, USA, 2019; pp. 350–354. [Google Scholar] [CrossRef]
  57. Cohn, R.; Holm, E. Unsupervised Machine Learning Via Transfer Learning and k-Means Clustering to Classify Materials Image Data. Integr. Mater. Manuf. Innov. 2021, 10, 231–244. [Google Scholar] [CrossRef]
  58. Rousseeuw, P.J. Silhouettes: A graphical aid to the interpretation and validation of cluster analysis. J. Comput. Appl. Math. 1987, 20, 53–65. [Google Scholar] [CrossRef]
  59. Schmid, P.J. Dynamic Mode Decomposition and Its Variants. Annu. Rev. Fluid Mech. 2022, 54, 225–254. [Google Scholar] [CrossRef]
Figure 1. HDMDc block scheme.
Figure 1. HDMDc block scheme.
Electronics 13 03047 g001
Figure 2. SRU line working scheme.
Figure 2. SRU line working scheme.
Electronics 13 03047 g002
Figure 3. Schematic representation of the debutanizer column (DC) with indication of the location of the hardware measuring devices, the model exogenous input, u, and soft sensor model output, y.
Figure 3. Schematic representation of the debutanizer column (DC) with indication of the location of the hardware measuring devices, the model exogenous input, u, and soft sensor model output, y.
Electronics 13 03047 g003
Figure 4. SRU case study: percentage performance improvement P I % for (a) M A P E % and (b) R 2 at each prediction step, varying the input delay shifts, q u , in the MSA-HDMDc algorithm. The P I % was calculated for each of the identified models with respect to the baseline model with q u = q = 40 .
Figure 4. SRU case study: percentage performance improvement P I % for (a) M A P E % and (b) R 2 at each prediction step, varying the input delay shifts, q u , in the MSA-HDMDc algorithm. The P I % was calculated for each of the identified models with respect to the baseline model with q u = q = 40 .
Electronics 13 03047 g004
Figure 5. SRU case study: MSA model performances: (a) M A P E % , (b) R 2 for ARX, FIR, and MSA-HDMDc models by varying the reduced order, p, of the Ω matrix in the p r a n g e { 201 , 202 , 210 , 220 , 240 } and considering the matrix X H at full-order r m a x = 40 .
Figure 5. SRU case study: MSA model performances: (a) M A P E % , (b) R 2 for ARX, FIR, and MSA-HDMDc models by varying the reduced order, p, of the Ω matrix in the p r a n g e { 201 , 202 , 210 , 220 , 240 } and considering the matrix X H at full-order r m a x = 40 .
Electronics 13 03047 g005
Figure 6. SRU case study: barplot of P I % for (a) M A P E % and (b) R 2 with Ω matrix order reduction p o p t = 202 and varying the X matrix reduction order in the range r r a n g e { 18 , 20 , 23 , 25 , 30 , 35 } . The P I % was calculated for each of the identified models with respect to the reference model with p o p t = 202 and r = r m a x = 40 .
Figure 6. SRU case study: barplot of P I % for (a) M A P E % and (b) R 2 with Ω matrix order reduction p o p t = 202 and varying the X matrix reduction order in the range r r a n g e { 18 , 20 , 23 , 25 , 30 , 35 } . The P I % was calculated for each of the identified models with respect to the reference model with p o p t = 202 and r = r m a x = 40 .
Electronics 13 03047 g006
Figure 7. SRU case study: regression plots of predicted output at 30 steps versus the target measured output, y 1 : (a) ARX model, (b) FIR model, (c) MSA-HDMDc model with optimal parameters q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 .
Figure 7. SRU case study: regression plots of predicted output at 30 steps versus the target measured output, y 1 : (a) ARX model, (b) FIR model, (c) MSA-HDMDc model with optimal parameters q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 .
Electronics 13 03047 g007
Figure 8. SRU case study: comparison of the measured output ( y 1 ) with the predicted ones at 30-step-ahead for the baseline and the MSA-HDMDc models with optimal parameters q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 .
Figure 8. SRU case study: comparison of the measured output ( y 1 ) with the predicted ones at 30-step-ahead for the baseline and the MSA-HDMDc models with optimal parameters q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 .
Electronics 13 03047 g008
Figure 9. SRU case study: analysis of M A P E % computed using time batches of 100 samples for a 30-step-ahead prediction on a selected interval of the test dataset. The corresponding normalized input signals and associated clusters are also included. 1st panel: time evolution of the inputs, 2nd panel: input clusters, 3rd panel: time evolution of M A P E % .
Figure 9. SRU case study: analysis of M A P E % computed using time batches of 100 samples for a 30-step-ahead prediction on a selected interval of the test dataset. The corresponding normalized input signals and associated clusters are also included. 1st panel: time evolution of the inputs, 2nd panel: input clusters, 3rd panel: time evolution of M A P E % .
Electronics 13 03047 g009
Figure 10. DC case study: MSA model performances in terms of (a) M A P E % , (b) R 2 for ARX, FIR and MSA-HDMDc models by varying the reduced order, p, of the Ω matrix in the p r a n g e { 65 , 66 , 70 , 84 } and considering the matrix X H at full-order r m a x = 12 .
Figure 10. DC case study: MSA model performances in terms of (a) M A P E % , (b) R 2 for ARX, FIR and MSA-HDMDc models by varying the reduced order, p, of the Ω matrix in the p r a n g e { 65 , 66 , 70 , 84 } and considering the matrix X H at full-order r m a x = 12 .
Electronics 13 03047 g010
Figure 11. DC case study: barplot of P I % for (a) M A P E % and (b) R 2 with Ω matrix order reduction p o p t = 66 and varying the X matrix reduction order in the range r r a n g e { 4 , 5 , 8 , 12 } . The P I % was calculated for each of the identified models with respect to the reference MSA-HDMDc model with p o p t = 66 and r = r m a x = 12 .
Figure 11. DC case study: barplot of P I % for (a) M A P E % and (b) R 2 with Ω matrix order reduction p o p t = 66 and varying the X matrix reduction order in the range r r a n g e { 4 , 5 , 8 , 12 } . The P I % was calculated for each of the identified models with respect to the reference MSA-HDMDc model with p o p t = 66 and r = r m a x = 12 .
Electronics 13 03047 g011
Figure 12. DC case study: comparison of the measured output (y) with the predicted one at (a) 5-step-ahead (30 min) and (b) 10-step-ahead (60 min) for the baseline and the MSA-HDMDc models with the optimal parameters q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 on a selected interval of the test dataset.
Figure 12. DC case study: comparison of the measured output (y) with the predicted one at (a) 5-step-ahead (30 min) and (b) 10-step-ahead (60 min) for the baseline and the MSA-HDMDc models with the optimal parameters q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 on a selected interval of the test dataset.
Electronics 13 03047 g012aElectronics 13 03047 g012b
Figure 13. DC case study: analysis of M A P E % computed using time batches of 100 samples for a 5-step-ahead prediction on a selected interval of the test dataset. The corresponding normalized input signals and associated clusters are also included. 1st panel: time evolution of the inputs, 2nd panel: input clusters, 3rd panel: time evolution of M A P E % .
Figure 13. DC case study: analysis of M A P E % computed using time batches of 100 samples for a 5-step-ahead prediction on a selected interval of the test dataset. The corresponding normalized input signals and associated clusters are also included. 1st panel: time evolution of the inputs, 2nd panel: input clusters, 3rd panel: time evolution of M A P E % .
Electronics 13 03047 g013
Table 1. Input and output variables of the SRU models.
Table 1. Input and output variables of the SRU models.
VariableDescription
u ( 1 ) = M E A _ G A S gas flow in the MEA chamber (NM3/h)
u ( 2 ) = A I R _ M E A airflow in the MEA chamber (NM3/h)
u ( 3 ) = S W S _ G A S total gas flow in the SWS chamber (NM3/h)
u ( 4 ) = A I R _ S W S total airflow in the SWS chamber (NM3/h)
u ( 5 ) = A I R _ M E A 2 secondary air flow (NM3/h)
y 1 = [ H 2 S ] H 2 S concentration (output 1) (mol%)
y 2 = [ S O 2 ] S O 2 concentration (output 2) (mol%)
Table 2. Input and output variables of the DC models.
Table 2. Input and output variables of the DC models.
VariableDescription
u ( 1 ) = T 040 top temperature (°C)
u ( 2 ) = P 011 top pressure (Kg/cm2)
u ( 3 ) = F 015 top reflux (m3/h)
u ( 4 ) = F 018 top flow (m3/h)
u ( 5 ) = T 004 side temperature (°C)
u ( 6 ) = ( T 036 + T 037 ) / 2 T 036 and T 037 bottom temperatures (°C)
y = F C 4 C4 concentration in the bottom flow (%)
Table 3. SRU case study: performance comparison for the selection of the q o p t . The mean value over 20 subsets of data of the M A P E % is reported for different state time shifts, q. The KPI is evaluated for a 30-step-ahead prediction. The P I % is reported considering q = 40 as the reference value.
Table 3. SRU case study: performance comparison for the selection of the q o p t . The mean value over 20 subsets of data of the M A P E % is reported for different state time shifts, q. The KPI is evaluated for a 30-step-ahead prediction. The P I % is reported considering q = 40 as the reference value.
State Time-Shift Optimization
q 2030405060
MAPE % 5.53 5.22 5.19 5.28 5.46
PI % 6.66 0.58 0 1.79 5.13
Table 4. SRU case study: M A P E % values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 ).
Table 4. SRU case study: M A P E % values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 ).
MAPE %
Steps 1 5 10 15 20 25 30
ARX 2 . 75 8.0810.9112.7013.7414.2614.55
FIR11.8411.8411.8411.8411.8411.8411.84
MSA-HDMDc5.56 5 . 85 6 . 24 6 . 59 6 . 90 7 . 15 7 . 35
Bold values represent the best performance for the specific prediction steps column.
Table 5. SRU case study: R 2 values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 ).
Table 5. SRU case study: R 2 values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 40 , q u o p t = 40 , p o p t = 202 , r o p t = 25 ).
R 2
Steps 1 5 10 15 20 25 30
ARX0.950.490.18−0.02−0.15−0.21−0.24
FIR0.260.260.260.260.260.260.26
MSA-HDMDc0.730.700.670.640.620.610.59
Bold values represent the best performance for the specific prediction steps column.
Table 6. DC case study: performance comparison for the selection of the q o p t . The mean value over 70 subsets of 100 of data samples of the M A P E % is reported for different state time shifts, q. The KPI is evaluated for a 20-step-ahead prediction. The P I % is reported considering q = 12 as the reference value.
Table 6. DC case study: performance comparison for the selection of the q o p t . The mean value over 70 subsets of 100 of data samples of the M A P E % is reported for different state time shifts, q. The KPI is evaluated for a 20-step-ahead prediction. The P I % is reported considering q = 12 as the reference value.
State Time-Shift Optimization
q 101215172030
MAPE % 25.73 24.87 25.32 26.43 27.52 30.19
PI % 3.43 0 1.81 6.28 10.65 17.20
Table 7. DC case study: M A P E % values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 ) in the test dataset.
Table 7. DC case study: M A P E % values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 ) in the test dataset.
MAPE %
Steps 2 5 10 20
ARX4.4413.7529.4053.91
FIR57.4257.4257.4257.42
MSA-HDMDc 1 . 66 6 . 01 12 . 93 23 . 13
Bold values represent the best performance for the specific prediction steps column.
Table 8. DC case study: R 2 values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 ) in the test dataset.
Table 8. DC case study: R 2 values at different prediction steps obtained for the considered models: ARX, FIR, MSA-HDMDc ( q o p t = 12 , q u o p t = 12 , p o p t = 66 , r o p t = 5 ) in the test dataset.
R 2
Steps 2 5 10 20
ARX0.9830.8540.347−1.20
FIR−0.99−0.99−0.99−0.99
MSA-HDMDc 0 . 998 0 . 974 0 . 890 0 . 661
Bold values represent the best performance for the specific prediction steps column.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Patanè, L.; Sapuppo, F.; Xibilia, M.G. Soft Sensors for Industrial Processes Using Multi-Step-Ahead Hankel Dynamic Mode Decomposition with Control. Electronics 2024, 13, 3047. https://doi.org/10.3390/electronics13153047

AMA Style

Patanè L, Sapuppo F, Xibilia MG. Soft Sensors for Industrial Processes Using Multi-Step-Ahead Hankel Dynamic Mode Decomposition with Control. Electronics. 2024; 13(15):3047. https://doi.org/10.3390/electronics13153047

Chicago/Turabian Style

Patanè, Luca, Francesca Sapuppo, and Maria Gabriella Xibilia. 2024. "Soft Sensors for Industrial Processes Using Multi-Step-Ahead Hankel Dynamic Mode Decomposition with Control" Electronics 13, no. 15: 3047. https://doi.org/10.3390/electronics13153047

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop